Download as pdf or txt
Download as pdf or txt
You are on page 1of 963

electromagnetic waves series 28

The
Handbook of
Antenna
Design
VOLUME 2
Editors
A.W. Rudge  K. Milne
A.D. OIver  P. Knight
Peter Peregrinus Ltd on behalf of the Institution of Electrical Engineers
IEE ELECTROMAGNETIC WAVES SERIES 16
SERIES EDITORS: PROFESSOR P.J.B. CLARRICOATS
E.D.R. SHEARMAN
AND J.R. WAIT

The
Handbook of

VOLUME 2
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
The
Handbook of
uvJLK.

VOLUME 2
Editors
A.W.Rudge K. Milne
A.D.OIver P.Knight
Peter Peregrinus Ltd on behalf of the Institution of Electrical Engineers
List of Contributors

VOLUME 2

EDITORS CHAPTER LEADERS CONTRIBUTORS

A.W. Rudge R.C. Hansen K.G. Corless


K. Milne G.. Borgiotti D. Hicks
A.D. Olver D.E.N. Davies
P. Knight R.H.J. Cary
J. Belrose
M. Radford
A. Burberry
W.T. Blackband

Published by Peter Peregrinus Ltd., London, UK.

©1983: 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.

British Library Cataloguing in Publication Data

The Handbook of antenna design.


Vol. 2. —(Electromagnetic waves; 16)
1. Antennas (Electronics)—Design and construction
I. Rudge, A.W. I I , Series
621.384V35 TK7871.6

ISBN 0-906048-87-7

Printed in England by Short Run Press Ltd., Exeter


Contents

List of contributors iv
Preface xiii
9 Linear arrays - R. C. Hansen 1
Introduction 1
9.1 Design of array excitation 2
9.1.1 Narrow beam, low sidelobe pattern 4
9.1.2 Pattern synthesis 42
9.1.3 Superdirective arrays 54
9.1.4 Thinned arrays 68
9.1.5 Array tolerances and limits 73
9.2 Array feeds 79
9.2.1 Resonant arrays 81
9.2.2 Travelling-wave arrays 87
9.2.3 Corporate line and distributed feeds 98
9.2.4 Phaser and sub-array lobes 104
9.3 Array elements 109
9.3.1 Dipoles 109
9.3.2 Waveguide slots 113
9.3.3 Scan compensated elements 121
9.3.4 Printed-circuit antennas 126
9.3.5 Moderate gain elements 133
9.4 References 134

10 Planar arrays - R. C. Hansen 141


Introduction 141
10.1 Array excitation 141
10.1.1 Array lattice 141
10.1.2 Beamwidth 143
10.1.3 Grating lobes 147
10.1.4 Quantisation lobes 151
10.1.5 Directivity 151
10.1.6 Planar array excitation 156
10.2 Array feeds 161
10.2.1 Fixed beam arrays 161
10.2.2 Electronic scanning in one plane 166
10.2.3 Electronic scanning in both planes 167
10.3 Mutual coupling 170
10.3.1 Fundamentals 172
vi Contents

10.3.2 Element-by-element approaches 181


10.3.3 Periodic cell approach 203
10.3.4 Scan compensation 218
10.4 References 222

11 Conformal arrays - G. Borgiotti 227


11.1 Introduction 227
11.2 Conformal array analysis 229
11.2.1 General 229
11.2.2 Formal characterisation of an array 230
11.2.3 Internal and external mutual couplings 233
11.2.4 Multimode aperture elements: Generalised scattering coefficients 235
11.2.5 Mutual admittances and scattering coefficients 238
11.3 Cylindrical array of apertures 240
11.3.1 Field of a magnetic current on a circular conducting cylinder 240
11.3.2 Harmonic series expressions for the mutual and self admittances 242
11.3.3 GTD surface field calculation 244
11.3.4 GTD expressions for the mutual admittances 247
11.3.5 Radiation pattern evaluation 252
11.4 Cylindrical arrays with a large number of elements:
The periodic structure approach 257
11.4.1 Periodic infinite array model 257
11.4.2 Eigenexcitations and eigenpatterns and their use in array analysis 258
11.4.3 Evaluation of the eigenvalues of the admittance matrices 261
11.4.4 Array RGP for an arbitrary excitation: Single excited element RGP 264
11.4.5 Aperture matching and element pattern design 265
11.4.6 Numerical examples and discussion 267
11.5 GTD evaluation of mutual admittances:
Generalisation to a generic convex surface 270
11.6 Conical arrays 274
11.6.1 Self and mutual admittances 274
11.6.2 Radiation pattern 278
11.7 Acknowledgement 284
11.8 Appendices 284
11.8.1 Modal admittances for rectangular and circular waveguides 284
11.8.2 Fourier transforms of the vector mode functions for rectangular and circular
waveguide radiators 286
11.8.3 Fock functions 288
11.8.4 Evaluation of the radiation pattern of an aperture on a large conducting
cylinder 290
11.8.5 Eigenex citations of the periodic cylindrical array 295
11.9 References 296

12 Circular arrays - D. E. N. Davies 298


12.1 Introduction 298
12.2 Beam cophasal excitation 299
12.2.1 Wideband performance 302
12.2.2 Electronic scanning for beam-cophased patterns 302
12.3 Amplitude-mode and phase-mode excitation of circular arrays 305
12.3.1 Concept of phase and amplitude modes for continuous circular arrays 305
12.3.2 Phase modes in circular arrays employing a finite number of
omnidirectional elements 309
12.3.3 Phase modes in circular arrays employing directional elements 310
12.4 Excitation networks for phase modes on circular arrays 314
Contorts

12.5 Mutual coupling in circular arrays 318


12.6 Applications of circular arrays 319
12.6.1 Communication applications 319
12.6.2 Radar applications 321
12.6.3 Circular arrays in direction finding 323
12.7 References 327

13 Array signal processing - D. E. N. Davies, K. G. Corless, D. S. Hicks and K. Milne 330


13.1 Introduction 330
13.1.1 Applications of array signal processing 331
13.2 ASP receiving systems for radar 332
13.2.1 Multiple beams 332
13.2.2 Multiple beams from element sampling 336
13.2.3 Equivalence between multiple beams and Nyquist rate beam scanning 337
13.2.4 Sidelobe levels in floodlight transmission systems 341
13.2.5 Compatability of ASP with other radar processing schemes 341
13.3 ASP in combined transmit and receive antenna systems 342
13.3.1 Pulse compression in angle by within pulse scanning on transmission 343
13.3.2 Array thinning on both transmit and receive by the use of ASP 344
13.3.3 Multiple beam operation on transmit and receive 347
13.3.4 Within pulse time sampling of transmitting array 347
13.3.5 Within pulse frequency scanning 348
13.3.6 Directionally decorrelated transmitter beams 349
13.3.7 Alternative configurations including multistatic systems 353
13.3.8 Signal processing technologies 354
13.4 Multiplicative signal processing and post-demodulation synthetic aperture
techniques 357
13.4.1 Interferometers 357
13.4.2 Interferometers with multiplicative processing 359
13.4.3 Principles of multiplicative arrays 360
13.4.4 Spatial frequency response for multiplicative arrays 363
13.4.5 Multiple source excitation of multiplicative arrays 365
13.4.6 Signal and noise performance of multiplicative arrays 367
13.4.7 Post-demodulation synthetic aperture antennas 368
13.5 Coherent synthetic aperture antennas 371
13.5.1 Principle 371
13.5.2 Unfocused synthetic aperture 374
13.5.3 Focused synthetic aperture 376
13.5.4 Minimum number of samples 380
13.5.5 Signal-processing hardware 382
13.5.6 Radar PRF 382
13.5.7 Doppler-frequency interpretation 382
13.5.8 Radar performance 383
13.5.9 Response to target scintillation and motion 384
13.5.10 Applications 385
13.6 Null steering 385
13.6.1 Introduction 385
13.6.2 Multiple nulls 386
13.6.3 Multi-element null-steering arrays 388
13.6.4 An orthogonal beam-forming network 391
13.6.5 Multi-element array with single null steering 393
395
13.6.6 Sidelobe canceller
13.6.7 Discussion 397
13.7 Adaptive null steering 399
viii Contents

13.7.1 Introduction 399


400
13.7.2 Sidelobe cancellers
13.7.3 Adaptive arrays 407
454
13.8 References

14 Radomes - R. J. Carey 457


14.1 Introduction 457
14.1.1 Definitions 457
14.1.2 History and background 457
14.1.3 Constructions 458
14.1.4 Disadvantages and advantages of radomes 460
14.2 Basic radome requirements 461
14.2.1 General requirements 461
14.2.2 Environmental requirements 461
14.3 Basic electrical properties of dielectric radome construction 462
14.3.1 Transmission parameters 462
14.3.2 Transmission of a single-layer construction 463
14.3.3 Transmission of an A sandwich construction 464
14.3.4 Transmission of the B sandwich construction 466
14.3.5 Transmission of the C sandwich and further multi-layer constructions 468
14.4 Design of shaped radomes 469
14.5 Radome materials 476
14.5.1 Materials and their requirements 476
14.5.2 Fibre-reinforced resin laminates 478
14.5.3 Core materials for resin-fibre laminates 433
14.5.4 High-temperature materials 488
14.5.5 Loaded and artificial dielectrics 498
14.5.6 Rain-erosion-resistant materials 499
14.6 Radome structures 501
14.6.1 Near spherical rigid shapes (usually on land or sea) 501
14.6.2 Non-spherical rigid shapes (airborne applications) 503
14.7 Airborne radomes 507
14.7.1 Wide-band transmission designs 507
14.7.2 Multi-band radome designs 511
14.7.3 Aberration and its reduction 515
14.7.4 Cross-polarisation introduced by radomes, and its reduction 521
14.7.5 Reduction of radome reflections 528
14.7.6 Effects of radome hardware (rain-erosion cap, pitot systems and
lightning protection) 533
14.7.7 Prediction and reduction of radome hardware degradation 536
14.8 Ground and ship borne radomes 536
14.8.1 Environmental considerations for ground radomes 537
14.8.2 Air-supported radomes 541
14.8.3 Rigid radomes 542
14.9 Further radome developments 549
14.10 Acknowledgements 5 50
14.11 References 5 50

15 VLF, LF and MF Antennas - J. Belrose 553


15.1 Introduction 553
15.2 Analysis 554
15.2.1 Characteristics of vertical antennas 554
15.2.2 Parameters needed to specify performance 556
15.2.3 Methods for calculating antenna parameters 557
Contents ix

15.2.4 Radiation resistance 559


15.2.5 Antenna impedance 562
15.2.6 Relation between electrical height and physical height 564
15.2.7 Mutual impedance between monopole antennas 565
15.2.8 Effect of ground conductivity and ground systems on antenna
terminal impedance 569
15.2.9 Effect of the finite conductivity of the earth on the vertical antenna
radiation pattern 574
15.2.10 Conductor and tuning-coil losses 577
15.3 Practical antenna design: Loaded antennas 578
15.3.1 Types and methods of feed 578
15.3.2 T-type antenna 580
15.3.3 Transmission-line radiator 584
15.3.4 Diamond antenna 586
15.3.5 Large VLF antenna 587
15.3.6 Umbrella antenna 588
15.3.7 Multiple-tuned VLF antennas 597
15.4 Practical antenna design: Unloaded antennas 598
15.4.1 Vertical monopolesh < \/4 598
15.4.2 Vertical monopoles \/4< h < 5/8\ 599
15.4.3 Vertical monopoles h < 5/8\ 603
15.4.4 Folded monopoles 604
15.5 Broadcast antennas 606
15.5.1 Introduction 606
15.5.2 LF broadcast antennas 607
15.5.3 Anti-fade antennas 608
15.5.4 Directional arrays 611
15.5.5 Distortion of MF broadcasting antenna patterns due to re-radiation 615
15.6 Portable and mobile antennas 624
15.6.1 VLF/LF aircraft wire antennas 625
15.6.2 Dipole antennas on the surface of the ground 625
15.6.3 Electrically short mobile whip antennas 627
15.6.4 Interaction of antennas with their platforms 629
15.6.5 Tuning and matching electrically short antennas 631
15.6.6 Ground effects when only a few radial wires are employed 633
15.7 Long wires antennas, dipoles and loops 634
15.7.1 Travelling-wave antennas 634
634
15.7.2 Inverted-V dipole antenna
636
15.7.3 Half-delta loop antenna
636
15.8 Receiving antennas
15.9 Practical antenna design 640
15.9.1 Guyed radio antenna masts 640
15.9.2 Base insulators for guyed masts 640
15.9.3 Guy strain insulators 643
15.9.4 Top loading design 644
15.9.5 Antenna tuning unit 648
15.9.6 Arrangement of the antenna tuning unit 651
15.9.7 Feeding a single tower radiator at two MF frequencies 652
15.9.8 Unique installations 655
15.10 References 658

16 High-frequency antennas - M. F. Radford 663


16.1 Introduction to HF 663
663
16.1.1 HF propagation
x Contents

16.1.2 Characteristics of HF signals 669


16.2 The ground at HF 672
16.2.1 Ground reflections 672
16.2.2 Ground improvement 675
16.3 Non-resonant HF antennas 677
16.3.1 Long-wire antenna 677
16.3.2 Vee antennas 679
16.3.3 Rhombic antenna 680
16.3.4 Beverage antenna 682
16.4 Resonant HF antennas 683
16.4.1 Monopole antenna 683
16.4.2 Dipoles and slot antennas 686
16.4.3 Loop antennas 687
16.5 Directional HF arrays 689
16.5.1 Endfire arrays 689
16.5.2 Broadside arrays 691
16.5.3 Circular arrays 697
16.6 Logarithmically periodic HF antennas 699
16.6.1 Early log-periodic antennas 699
16.6.2 Logarithmic dipole antennas 700
16.6.3 Logarithmic antenna arrays 706
16.7 Unorthodox HF antennas 708
16.7.1 HF horn antenna 708
16.7.2 HF wire and lens antenna 709
16.8 HF antennas for mobile applications 711
16.8.1 HF antennas for land vehicles 711
16.8.2 HF antennas for aircraft 713
16.8.3 HF antennas for ships 716
16.9 HF antenna design 717
16.9.1 Theoretical design techniques 717
16.9.2 Scale-modelling techniques 717
16.9.3 HF antenna performance measurement 719
16.9.4 Practical HF antenna design 720
16.9.5 HF antenna siting 722
16.10 References 723

17 VHF and UHF antennas - A. Burberry 725


17.1 Introduction 725
17.1.1 Definition of frequency range 725
17.1.2 Applications 725
17.1.3 Classes of antenna 726
17.2 Omnidirectional antennas 726
17.2.1 Single elements 727
17.2.2 Stacked elements 749
17.2.3 Effects of mounting structure 753
17.3 Directional antennas 753
17.3.1 Aperiodic reflectors 754
17.3.2 Parasitic elements 762
17.3.3 Log periodic antennas 767
17.3.4 Helical antenna 770
17.3.5 Slotted cylinder 772
17.3.6 Combination of elements 773
17.3.7 Significant design parameters 775
17.4 Vehicular antennas 776
Contents xi

11 A.I Antenna siting methods 776


17.4.2 Influence of finite conducting surfaces 781
17.4.3 Vehicular constraints 787
17.4.4 Land-based vehicles 788
17.4.5 Marine vehicles 793
17.4.6 Aircraft 794
17.4.7 Missiles and rockets 819
17.4.8 Satellites 821
17.4.9 Antennas on persons and animals 823
17.4.10 Methods of measurement on vehicles 827
17.5 Position-finding antennas 830
17.5.1 Homing systems 830
17.5.2 Tracking and direction-finding systems 831
17.5.3 Transmitting antennas for navigation systems 837
17.6 Dielectric-loaded antennas 840
17.6.1 Dielectric loading 840
17.6.2 Dielectric covers 842
17.7 Circularly polarised antennas 845
17.7.1 Definition 845
17.7.2 Combination of elements 846
17.8 Multiple installations 850
17.8.1 Effect of one antenna on another 850
17.8.2 Multiband antennas 854
17.9 References 856

18 Coaxial transmission lines and components - A. Blackband 860


18.1 Introduction 860
18.2 Coaxial transmission lines 860
18.2.1 Frequency limitations 867
18.2.2 Voltage and thermal ratings 867
18.2.3 Screened twin lines 867
18.3 Open-wire transmission lines 871
18.4 Miscellaneous parallel transmission lines 880
18.4.1 Shapes of conductor cross-sections 881
18.4.2 Conductors in enclosures 883
18.4.3 Lines with strip or ribbon conductors 883
18.5 Conical transmission lines 887
18.6 Calculation of characteristic impedances 902
18.6.1 Constants in the expressions for Zo 902
18.6.2 Range of validity for approximate formulae for Zo 902
18.6.3 Methods of calculating or estimating the Zo of other lines 904
18.7 Layout of cable runs for minimum interference 906
18.7.1 Coupling between coaxial cables 906
18.7.2 Mutual screening of vulnerable screened circuits 912
18.8 Junctions in coaxial line systems 914
18.8.1 Discontinuities at junctions 914
18.8.2 Impedance-matching sections 916
18.9 Balancing devices or baluns 917
18.10 Power dividers 922
18.10.1 The'rat race' 922
18.10.2 The Bagley polygon 923
18.11 Plugs and sockets or connectors 925
18.12 References 928
Index 930
Preface

The Handbook of Antenna Design was conceived at a discussion between two of


the editors following a meeting of the Professional Group on Antennas and Propa-
gation at the Institution of Electrical Engineers in London during 1977. The
original concept was to produce a text which dealt with the principles and appli-
cations of antenna design with particular emphasis upon developments which had
occurred in the previous decade. The authors of the text were to be invited from a
select list of internationally respected experts in the field, to provide a compre-
hensive treatise on the design of antennas in the frequency bands from LF to
microwaves.
In the event the sheer magnitude of the task has imposed its own limitations.
Many topics which are worthy of treatment have been omitted and although the
coverage of this two-volume text is undoubtedly broad, the editors are fully
conscious of the fact that additional volumes would not be difficult to fill. To
condense the material to the greatest extent possible the contributing authors
were asked to include detailed theoretical developments only where these were
fundamental to a proper understanding of the design principles, and were not
readily available in existing well-established texts. Although much emphasis has
been placed upon the provision of formuli and data which is applicable to the
design process, it was recognised from the outset that an adequate understanding
of the basic principles involved is essential for good design practice and in encour-
aging innovation.
The extensive efforts which have been devoted to the development of antennas
and the associated design and measurement techniques over the past few decades
is, perhaps, not surprising. The electromagnetic spectrum is a limited resource
and with the very rapid growth of system applications in communications, navi-
gation and radar, there has been a continuous pressure for more effective and
efficient transmission and reception of electromagnetic radiation. The crowded
spectrum has led to more demanding specifications upon the performance of
radiating devices and has thereby created an underlying need for improved design
techniques.
The need for more accurate design and the availability of powerful digital com-
puters have combined to guide antenna design along a path which is providing a
continuous transition from empirical art to mathematical science. This trend has
xiv Preface

been nowhere more evident that at the shorter wavelengths, where the combination
of digital computer and mathematical method has found numerous applications in
the design of quasi-optical antennas. Computer codes based upon geometric optics,
physical optics, the geometric theory of diffraction, Fourier transform theory, arid
spherical-wave analysis now provide essential tools in modern antenna design.
Thoughtful application of these analytical methods has lead, not only to improved
computations of antenna radiation characteristics, but also to improved insight into
the underlying physical phenomena.
The first volume of this text provides the mathematical background and a large
number of examples of the use of mathematical methods in the design of reflector
and lens antennas, including their primary-feeds. Quasi-optical antennas, including
hybrid configurations, where the reflector or lens is used in conjunction with a
complex array feed, are particularly well suited to the use of computer-aided design
methods and major performance improvements can be realised. These topics are
dealt with in Chapters 2-6 of the first volume.
Low and medium gain microwave antennas have many applications, of which
serving as a primary-feed is only one. Chapter 7 provides design data for a diverse
range of radiators including helices and spirals, slots, microstrip antennas, backfire
antennas, dielectric rods and horns.
Advances in antenna performance must be matched by improved electrical
measurements and there have been a number of significant developments in antenna
metrology during the past decade. The key developments in this important area are
described in the last chapter of the first volume.
The fundamental principles and design of antenna arrays has been the focus of a
great deal of attention during the past decade. Although cost still remains a barrier
in the wide scale deployment of large phased-arrays a continued growth in the
applications of antenna arrays can be anticipated. The importance of array tech-
nology has been recognised here in the dedication of the first five chapters of the
second volume to this theme. The subject matter covered includes linear, planar,
conformal and circular geometries in addition to array signal processing. It is
evident that the combination of antenna arrays with integrated-circuit technology
will have much to offer in the future in terms of both performance flexibility and
lower cost.
In practice many antennas are operated either behind or within a radome cover.
In such cases the electromagnetic characteristics of the radome must be considered
as a factor in arriving at the overall system performance. In Chapter 14 this
important topic is considered and design data are provided for a variety of radome
types.
The last few chapters of the second volume are concerned with the design of
antennas and coaxial components in the LF to UHF frequency bands. In general
design data for these bands are less well documented than their counterparts in
the microwave region, and the opportunity has been taken here to bring together
design information from a number of sources including a rather sparse and widely
dispersed literature.
Preface xv

In preparing these volumes the editors are indebted to the authors, all of whom
are well known experts in their fields, who have somehow found time in their very
busy schedules to organise and prepare their contributions. The authors have
exhibited a remarkable degree of patience in seeing through a project which has
taken several years to complete. The work described here has been performed by a
great many engineers and scientists over a period of many years and although a
sensible effort has been made to ensure that the references and credits are accurate,
the editors offer apologies for any errors or omissions in this respect.
For our part as editors we have sought to bring together the contributions into
a cohesive whole and have on occasions made modifications or deletions to achieve
this end. Any resulting omissions, misprints or errors in the text remains our
responsibility and we shall be glad to have notice of them to correct any future
editions. Although some effort has been made to standardise the notations
employed throughout the text, the two volumes represent the efforts and individual
styles of 28 contributing authors and, in view of the excessive work entailed, a
comprehensive cross-correlation between chapters has not been undertaken. Vector
notation^ have been indicated by the use of either bold type as in E, H, or by use of
bars as E, H. References, equations and figures are numbered in sequence on a
chapter basis, and the references are listed at the end of each chapter. A compre-
hensive index is provided at the end of each volume.
It is the editors hope that this handbook will be of value both to practising
design engineers and to students of antenna theory. In the preparation of the text
it has been assumed that the reader has at least a working knowledge of electro-
magnetic theory and antenna technology, however, the material provided should
be sufficiently detailed to guide the reader from this stage. The review of the basic
properties of antennas contained in Chapter 1 provides an indication of the
technical level required and includes appropriate references to further background
reading.
Finally the editors wish to acknowledge with thanks the many organisations and
institutions for their kind permission to use figures, photographs and other
information.
Chapter 9

Linear arrays
R. C. Hansen

Introduction

Linear arrays of radiators, electromagnetic and others, have a history of develop-


ment over many decades. In recent years many developments have come together
to disclose a many faceted understanding, along with a computer-based design
capability. Neither are yet complete, but linear arrays have achieved a significant
level of maturity. Key advances are the study and placement of space-factor zeros,
Hermitian quadratic-form constrained synthesis, low sidelobe designs, calculation of
element admittances, mutual coupling analysis including multi-mode scan compen-
sation, and finally computer design of arrays. As mentioned, the linear-array tech-
nology base is old; thus this chapter is necessarily eclectic. In many cases, however,
older data have been re-calculated using the current excellent computing capability
available almost everywhere, and new data has been calculated to supplement, and
to provide, a more complete picture.
The first and largest Section is concerned with design of the array excitation: the
set of complex numbers representing drive currents or voltages. Narrow-beam and
shaped-beam design principles are covered in detail for uniformly spaced arrays,
followed by constrained synthesis. The important topic of superdirectivity (super
gain) is treated in depth. Non-uniformly spaced (thinned) arrays, and array toleran-
ces are also treated. Section 2 covers array feeds: resonant (standing wave), travell-
ing wave, corporate, line and distributed. Salient lobes produced by phaser quantis-
ation or by sub-arraying are then studied. In Section 3 array elements are the topic,
primarily low gain elements such as slots, dipoles, microstrip elements etc.
Moderate gain elements such as spirals, log-periodics, horns etc. are covered in other
Chapters. Mutual coupling is not discussed in this Chapter; an extensive discussion
of mutual coupling, active element patterns, active impedance, element efficiency
etc. for both linear and planar arrays is given in Chapter 10.
It is a pleasure to thank many colleagues for help and counsel. This Chapter is
dedicated to the late Dr. Henry Jasik, a friend and major contributor to the antenna
art.
2 Linear arrays

9.1 Design of array excitation

The radiated fields from a linear array are a superposition (sum) of the fields
radiated by each element (in the presence of the other elements). The field of an
element is proportional to the excitation parameter, which is current for a dipole,
voltage for a slot, and mode voltage for a multiple-mode element. Self and mutual
impedance (or admittance) relates the applied voltage (current) to the excitation
current (voltage). This Section is concerned with the design of the set of excita-
tions, a complex number for each element.
A common notation will be used throughout this Chapter; it is that most often
used by array papers. The excitation vector will consist of an amplitude and phase
at each element. This discrete distribution is often called an aperture distribution,
where the discrete array is the aperture. The far-field radiation pattern, also called
the space factor, is just the Discrete Fourier Transform of the array excitation.
When the array is assumed to have isotropic elements, this DFT is the sum of exci-
tation coefficients times phase-path exponentials. The overall pattern is the product
of the isolated element pattern and the isotropic array space factor. This is the
'forced excitation' problem: the drive voltages are individually adjusted so that each
excitation current (using dipoles as an example) is exactly the value desired. In
practice many feeds produce a 'free excitation', that is the drive voltages are all
fixed and the excitation currents are those allowed by the 'active' element
impedance. Active implies a variation with observation angle 6. Ways of calculating
active array patterns using active impedances or active element patterns are discus-
sed in Section 9.4 and in Chapter 10. Returning to notation, the element spacing is d
where each d may be different, angles are measured from broadside, X is wave-
length, and the convenient array variable is u:
d n
u = - sin d
A

n = 1 2 3 L . . . N

Fig. 9.1 Linear array geometry

Uniform (equal) spacing is assumed in what follows. If the co-ordinate system starts
at one end of the array, as in Fig. 9.1, the space factor is, for an N element array:
iV
F(u) = I An exp; 2n(n - l)u (9.1)
Linear arrays 3

Again forced excitation is assumed, and An is the complex excitation, which may
include phasing. Later the phasing terms will be written explicitly for narrow-beam
and for shaped-beam patterns. Also it will be convenient later to move the co-
ordinate centre to the array centre.
The unit circle approach of Schelkunoff135 replaces the exponential by:
w = exp (/ 2m)

Then the pattern becomes a polynomial in w:

F(u) = £ Anwn ' 1 (9.2)

Real (physical) space corresponds to a portion of the unit circle in the complex w
plane. The polar angle in the w plane is \jj — 2mi, and the portion of the unit circle
covered is for:

-kd<\p<kd (9.3)
Thus half-wave spacing allows w to traverse the unit circle once, wavelength spacing
twice, etc. The w polynomial has N— 1 roots, but they do not have to lie on the
unit circle.

} main beam
w = 1 region

sidelobe
region

Fig. 9.2 Unit circle for 8-element array

The pattern is given by the product of the distances from the observation point
(w) to each of the zeros (see Fig. 9.2). As the point moves around the circle, lobes
build up then decay, and, as mentioned, zeros located on the circle give pattern
nulls. Zeros off the circle may give pattern minima. The uniform array is an
especially simple case in that:

F{u) = (9.4)
sin nit w-1
4 Linear arrays

Here the exponential term corresponds to the phase at the array centre and the
roots are equally spaced on the circle by 2n/N. This unit circle analysis is exactly
the same as the z-transform, which has gained some currency in circuit analysis in
recent years, z=w. Zero (root) placement will be discussed further in Section
9.1.1.3.

9.1.1 Narrow beam, low sidelobe patterns


A common pattern requirement is for high directivity and low sidelobes, useful for
radar, communications, mapping etc. Sidelobe level (SLL) is the amplitude of the
highest sidelobe, usually that closest to the main beam, normalised to the main
beam peak. It is convenient to also use sidelobe ratio (SLR) which is the inverse of
SLL. The main beam may be fixed at broadside or at some other angle. Or it may
be electronically scanned to any desired angle. The terms 'electronic scanning array'
and 'phased array' are synonomous. For maximum directivity* without super-
directivity, uniform excitation is used. When lower sidelobes than those provided
by uniform illumination are required, the aperture amplitude is tapered from the
centre to the ends. There are some general rules, derived by Taylor157 and others:
Symmetric amplitude distributions give lower sidelobes.
F(u) should be an even entire function of u.
A distribution with a pedestal produces a far out sidelobe envelope of \\u.
A distribution going linearly to zero at the ends produces a far out sidelobe
envelope of lju2.
A distribution that is non-zero at the ends (pedestal) is more efficient.
Zeros should be real (located on the w unit circle)
Far out zeros should be separated by unity (in u).
Although extensive theory has been developed for continous aperture distribu-
tions, there is little available for array distributions. For example, there are no dis-
crete distributions that yield a highly efficient tapered sidelobe pattern. Thus, in
designing most arrays, a continuous distribution is first constructed; this distribu-
tion is then sampled at the element locations to obtain the array excitations. This
process works well if the continuous distribution amplitude and phase are slowly
varying and without oscillatory components. In some cases sampling is not satisfac-
tory. A more sophisticated method matches zeros (Elliott46). First the continuous
distribution is selected, and the space factor is expressed as a product of zeros.
Since many continuous distributions contain trig or hyperbolic functions, these are
replaced by product representations which show zeros explicitly. Then the array
polynomial in w (eqn. 9.2) is equated, with the array-factor zeros made equal to the
space-factor zeros. The array factors are next multiplied out to get the array exci-
tations. Since the zeros of the pattern function control the spacing and amplitude
of all pattern lobes, equating zeros in general gives an array factor that is close

* This implies a linear phase progression over the array, corresponding to main beam angle 0O
Linear arrays 5
to the desired continuous space factor (pattern). If the calculated pattern is still not
sufficiently close to the desired pattern, a perturbation technique may be used. For
example, with a narrow beam pattern with a shaped sidelobe envelope where side-
lobe heights are slightly off, it may be assumed that the position of each sidelobe
peak is unchanged (valid since the peaks are rounded on top). Then each element
excitation may be perturbed and an error signal may be derived from the difference
between each lobe peak in the original pattern and in the pattern produced by the
perturbed excitations. Solution of the resulting equations yields values for the exci-
tation perturbations. Several iterations may be necessary for satisfactory pattern
adjustment.

9.1.1.1 Physics vs. mathematics: For uniform excitation, calculation of array


performance is relatively easy, and this is discussed in the next Section. Tapered
distributions used to reduce sidelobes are less easy. In the days BC*, aperture dis-
tributions were chosen for their easy integrability to closed-form solutions. For
example, a cosine to the nth power, on a pedestal, allows the sidelobe level to be
adjusted, and is readily calculated. There are, however, two disadvantages to these
distributions. In the example there are two parameters: power and pedestal height.
For given values of these, all important characteristics (beamwidth, aperture
efficiency, sidelobe level, beam efficiency etc.) can be calculated. But there is no
easy way to choose the two parameters to optimise efficiency for a given SLR, for
example. More fundamental is the second disadvantage: these distributions, depend-
ing on the values of parameters selected, may be quite inefficient in terms of beam-
width, aperture efficiency etc. Another example is the popular Gaussian. With a
pedestal (a truncated Gaussian) it is a multiple parameter distribution, and, what is
worse, the Fourier transform is complex. Thus in modern antenna work easy
mathematics has yielded to good physics: aperture distributions should be designed
by proper placement of pattern-function zeros, preferably using a single parameter.
This assures good physics which means highly efficient, low Q apertures. Design can
be handled via the computer. Taylor was a pioneer in this approach to narrow-
beamwidth low-sidelobe distributions; two of the most useful, and most used,
distributions were invented by him and will be discussed under Section 9.1.1.3.

9.1.1.2 Uniform ly excited arrays

Pattern: In writing the pattern of an array with equal-amplitude coefficients


(uniform excitation) it is convenient to sum the path lengths from each element to
the phase front starting at one end, and then multiplying by a phase factor repre-
senting the phase from the array end to the centre. This gives a real pattern and
works for the number of elements N both odd and even. With angles measured
from broadside, the main beam at 60, and element spacing (2, the pattern is:

Before computers
Linear arrays
sin Nrni
F(u) = f (9.5)
sin nu
where now
u = -(sin0 — sin0o)
A

The interelement phase shift is kd sin0o. When this is zero the beam is broadside.
By varying the interelement phase shift the beam position is controlled to produce
a main beam scan. Fig. 9.3 shows the variation in pattern that can be produced by
a two-element array by varying spacing and phase (Southworth141). Before taking
up the behaviour of these canonical uniform arrays when scanned, it is useful to
evaluate broadside performance.

no o Q
(0 (d) (e)

Fig. 9.3 Directivity patterns of 2-element array vs. spacing and phase ( T = 2-n) {courtesy:
Southworth)

Beamwidth: The half-power points on the beam of a uniform array are found
simply by putting sin Nm3/(N sin nu3) = \/o3. The solution of this for various N is
given in Fig. 9.4, where it may be seen that the variation inNu 3 is less than 1% for
N>1. In fact the error is only 5% for N= 3. For large arrays, the half-power
points can be written simply: Nu3 = ± 0.4429. For a beam scanned at angle 0O> this
gives the 3 dB beamwidth 03 (angle between the 3 dB points on the main beam as:
/ 0.443 X\ / 0.443 \\
0$ 2£ arc sin I sin d0 H — arc sin I sin 0o 1
y Nd ] \ Nd ]
or, for TV large,
0.866 X
(9.6)
Nd cos 0A
Linear arrays 7
It is useful here to consider how the beam changes with scan. At broadside the
beam is a rotationally symmetric disc (omitting element patterns) as sketched in
Fig. 9.5. As the beam scans toward endfire, the disc folds forward to make a cone.

0.52 i

0.46-

0.42 -

0.40
6 10 100
number of ele

Fig. 9.4 Normalised beam width

Fig. 9.5 Linear-array beam forming


a Disc at broadside
b Cone
c Pencil at endfire

A singular situation occurs when the 3 dB point gets to 90°. Beyond this scan angle
the beam has two peaks and the 'beamwidth' will double as the outside 3 dB points
are used. Finally at endfire a pencil beam results. Thus a linear array at broadside
yields directivity in one dimension, while at endfire it yields directivity in two
8 Linear arrays

dimensions. Of course the endfire beamwidth is broader. Returning to eqn. 9.6, at


broadside the beamwidth is
0.886 X
(9.7)
Nd

10 20 50 100

Fig. 9.6 Beam width broadening near endfire

Accuracy of this is better than 1% for Ndfk > 2. The beam collapse, where the 3 dB
point is at 90°, occurs for a scan angle of:
0.443X\
0 = arc sin II — (9.8)

and the endfire beamwidth is:


/ 0.886 X
03 ^ 2 (9.9)

and the accuracy of this is better than 1% for Nd/\ > 4. This beamwidth is larger
than the broadside value by 2.14 \/Nd/\. Thus for uniform linear arrays these
formulas are all valid forN> 20; for smaller arrays the value of Nu3 from Fig. 9.4 is
substituted in place of 0.443 in eqn. 9.6 and 03 can be determined for any scan
angle. Fig. 9.6 shows the beam broadening vs. angle. When the inner 3 dB pattern
point reaches 90°, the pattern cone is filling in and the usual definition of beam-
width is not useful.

Sidelobes: Sidelobes of the uniform array are well behaved; nulls of F(u) occur
at u — n/N, n= 1 to N — 1, and the peaks of F(u) occur at u that are solutions of
TVtanrrw = X^nN-nu. For large N, the equation becomes tan7V~7rw =Nnu, and the
Linear arrays 9

first solution is Nu = 1.4303. The sidelobe ratio for large arrays is the same as that
for uniform line sources: 13.26 dB, independent of main beam angle. For smaller
arrays the location (value of u) of the first sidelobe and the sidelobe ratio are shown
in Fig. 9.7 where it appears that arrays of less than 8 elements experience a signifi-
cant sidelobe-level degradation. Of course, the relation of sidelobe peak angles and
array length depends upon main beam position through u. The uniform linear array
has a sidelobe envelope that decays as nu approaches 7r/2, and as mentioned in Sec-
tion 9.1.1.3, this decay is important to maintain low aperture Q and tolerance in-
sensitivity.

47n 15

1 46-
sidelobe ratio, dB •

1 45- 13

12

• sidelobe position Nu

142 10

141

140
10 20 40 60 100

Fig. 9.7

Grating lobes: From the expression for the array pattern (eqn. 9.5) it is
obvious that maxima of unit value occur whenever u = «, an integer, because then
both sine terms are zero and the limit is unity. If d/X and 0O are chosen properly,
only one main beam exists in Visible' space, — 90° < 0 < 90°. Large spacings will
produce one or more additional main beams called grating lobes (GL) (see Fig. 9.8
for a sketch of a typical pattern). Physically the larger spacing allows the wave from
each array element to add in phase at the GL angle. From the condition given above,
u =n9 the location of grating lobes can be easily determined. The GL equation is
usually written as:
10 Linear arrays

(9.10)
X sin 60 — sin 6g

When the main beam is broadside, the first grating lobe is just appearing at 6g =
— 90°, when d = A. A symmetric GL appears at 8g — 90°. Beam scan to 0O = 90°
will produce a GL at —90° when d = A/2; this is the oft quoted rule that half-wave
spacing precludes grating lobes. Actually as d -> A/2 the GL approaches visible space,
with half of the GL visible at d = A/2. Higher-order grating lobes occur when n =
2, 3 or more. Radio-astronomy interferometers use spacings of several wavelengths
or more to get narrow beamwidth with few elements; many grating lobes result.
Fig. 9.9 shows the onset of the first grating lobe for various scan angles (main beam
angles), and GL angles vs. d/X. For example a 30° scan requires a spacing of 0.667A
or less to keep grating lobes at the edge of visible space or beyond.

-90 90

Fig. 9.8 Array pattern scanned to 45°


£=0.707
A.

Placement of GL null: It is often desirable to keep all of the grating lobe out of
visible space. One might in principle adjust d/X so the grating lobe at — TT/2 is equal
to the sidelobe envelope near there. Since the sides of the grating lobe are steep,
small changes in frequency or manufacturing tolerances may bring in a sizable por-
tion of the grating lobe. A better scheme puts the pattern zero adjacent to the grat-
ing lobe at — TT/2, SO that the entire grating lobe is comfortably excluded. This can
be done for uniformly illuminated arrays directly, but, as mentioned elsewhere,
there are no universal array distributions with adjustable sidelobe level and Q. Such
arrays are usually designed using continuous distributions such as the Taylor one-
parameter modified sinx/x line source. Thus this distribution will be used here to
find the reduction needed in d/X to put the grating lobe null at —nil* Unlike the
line source pattern, which has no grating lobes, the array pattern (for large d/X) is a
repetition of part of the line source pattern, as shown in Fig. 9.10. The sidelobes
This analysis is due to David Munger of Motorola, Phoenix, AZ, and Richard Phelan of Harris
Corp., Melbourne, Fl
Linear arrays 11
are lowest halfway between the main lobe and grating lobe. Using the Taylor one-
parameter distribution (see Section 9.1.1.3) the part of the pattern adjacent to the
main lobe is given by:
diiin^/u MJ Nd , ^N
F(«) = f
•z « = — (sm 0O ~ sin 6) (9.11)
TT\ U —B* A

where B is the parameter which is selected to determine the sidelobe level. The first
null occurs for ux — \/l + B\ and placing the grating lobe at — TT/2 gives:
Nd
ug = — (sin 0 O + 1) (9.12)
A

12-1
grating lobe angle

11-

10 20 30 40 50 60 80 90
main beam angle 90

Fig. 9.9 Element spacing vs. grating lobe angle

Using the grating lobe equation (eqn. 9.10), ug = N. Now the portion of the pattern
adjacent to the grating lobe is given by:
sin 7T\/(Ug — u)2 — B2
F(u) = (9.13)

The null adjacent to the grating lobe is at u = ug — uu and when it is now placed at
-TT/2:
12 Linear arrays

Nd . __ Nd,
8 l 8 (9.14)
X ° X
But from above ug — Ui=N — \ / l + B29 so the modified grating equation becomes:
d_ _ N-^/l+B2 i 1
(9.15)
X N sin 0O + 1

Fig. 9.10 Array pattern with grating lobe-,

Thus the array spacing reduction factor is:


N-\/l+B2
(9.16)
N
When the amplitude distribution is uniform, B = 0, and the reduction becomes
(N — 1)/JV, which is correct. Fig. 9.11 shows this spacing reduction factor vs.iV^ for
sidelobe ratios of 13.26, 20, 25, 30, 35 and 40dB. Reduction factors below 0.6 are
not plotted as they correspond to small values of iV: small arrays are not efficient in
producing low sidelobes.

Directional elements: Directional elements are often employed in linear arrays,


to reduce the wide-angle radiation in the plane of the array and/or to reduce the
beam width in the orthogonal plane.
Assuming that the radiation patterns of all elements are identical and are given
by H(6, 0), the overall pattern becomes
E(fi,4>) = F(u)H(6,<t>) (9.16a)
i.e. the overall pattern is the array factor F(u) multiplied by the element factor.

Directivity and effective aperture: The directivity of a linear array may be


found in the usual way by integrating the power radiation pattern over a sphere to
find the total radiated power and comparing this with the power density at the
angle of interest. The directivity, here denoted by G since losses are ignored, is
given by:
Linear arrays 13
2
G = o, <t>o)\

4n\H(do,<t>o)\2\F(uo)\2
(9.17)
Jo J-*a WW* <t>)\2\F(u)\2cosd dd d<t>
For isotropic elements Hid, 0) = I and the integration with respect to 0 may be
performed immediately since the array factor F(u) is rotationally symmetric (Fig.
9.12), giving:

G = (9.17a)
1
J-7T/2
cos Odd

10-

09-

08-

07-

SLR=13dB 20 25 30 35^0

0-5-

03
6 10 20 30 L0 60 100
number of element^ N

Fig. 9.11 Element spacing reduction factor

In the case of a uniform broadside array F(u) is symmetric and is given by equation
9.5, so that
N2
G = (9.17b)
cos0 dd
14 Linear arrays
Changing the variable to u gives:
N2d/X

r
G = (9.18)
sin2N7tu
2
sin 7rw
Jo
This can be integrated with the help of an expansion (Whittaker and Watson176):
I^d/X _ rd,
: £ (/V — «)cos 2nmt du
G ~ Jo n=l J
NdMN^\
__ — + — KT
2* {N — n) sine nkd (9.19)
A A n=l

Fig. 9.12 Line-source geometry

The directivity result, where sine is sinx/x, is:


N2
G = N-l (9.20)
N+2
n=l
Fig. 9.13 shows directivity for various arrays from 2 to 24 elements as a function of
d/X. The effect of the grating lobes can be seen around d = X. For larger spacings
the directivity increases until, around d = 2X, another drop represents the appear-
ance of second-order grating lobes. It is evident that whenever nkd = TT, the direc-
tivity is N:

i.e. G = for- = 0.5,1.0,1.5,


A

The peak directivity occurs just before the onset of the grating lobe and the minor
oscillations below the peak depend upon whether the boundary of real space
(sin# = 1) coincides with a peak sidelobe or with a zero. This behaviour can also be
interpreted as the effect of the mutual coupling between elements varying as the
spacing is varied. In fact, the summation in eq. 9.20 can be recognised as a weighted
Linear arrays 15
summation of the mutual resistances, since the normalised mutual resistance
between a pair of isotropic radiators spaced by 5 is simply sine (ks).
At this stage, it is worth recalling that the assumption underlying the curves shown
is that the currents in all elements have been forced to be equal. Achievement of

10 12

Fig. 9.13 Array-factor directivity (isotropic elements)

equal currents depends upon the design of the feeding network. Simply connecting
all the elements in parallel will not achieve equal currents, since the end elements
'see' a different mutual impedance than do the centre elements. The feeding network
has to allow for the effect of self-reactance and mutual reactance as well as the
mutual resistance of the elements.
16 Linear arrays
For a scan off broadside, the directivity result is:
N2
G = spi (9.21)
N + 2 £ (iV — «) sine wfaf cos («M sin 0O)

See also Section 9.1.3.3.


This can be used to calculate directivity at any scan angle, and will produce
directivity changes near grating lobe incidence similar to those of Fig. 9.13. Endfire
is a special case, with the result:

N
G = N.x (9.22)
1+- £(#-«)sine2n*rf
N n=i
This duplicates the broadside array results, but with spacing halved. Thus G = TV for
d/X = 0.25, 0.5, 0.75, . . . The beam at endfire as mentioned earlier is a pencil
beam, but the beamwidth is much broader than the broadside disc beam. Using the
general formula with half-wave spacing inserted, it is interesting to see how the
effective aperture (directivity divided by broadside directivity N) varies with scan
angle. Since the sine argument is nn, the directivity is N for any scan angle, and the
effective aperture is unity. This is in contrast to a uniformly excited continuous line
source (Hansen69) which has an endfire directivity of 4L/X, but a broadside direc-
tivity of 2Z//X. (Although the endfire 3 dB beamwidth is again much broader than
the broadside beamwidth, the total solid angle is smaller, cf. Fig. 9.5.)
For N large, maximum directivity occurs with a spacing of just under X for
broadside operation and just under X/2 for endfire operation, with a value
approaching 2N. Expressed in terms of aperture length L = (N—l)d, it is evident
that these maximum directivities approach the same values as those of the continu-
ous line source, viz. 2L/X at broadside and 4L/X at endfire.
Turning now to arrays of actual elements such as short dipoles or half-wave di-
poles (slots have similar patterns), the overall radiation pattern is given by the
element pattern multiplied by the array-factor pattern (isotropic elements), but the
overall directivity is not the product of the element and array directivities.
Parallel short dipoles provide directivity in the 0 dimension and the appropriate
element pattern to use in eqn. 9.17 is:
>) = V I ~ cosz0 sin> (9.23)
whilst collinear short dipoles provide directivity in the plane of the array with an
element pattern:

H(d,<t>) =eos(9 (9.23a)

The analysis proceeds along the same lines as before. The resulting directivity for a
uniform broadside array of parallel short dipoles is:
Linear arrays 17

G = 3N/2 ,923b)
. . 3*L* \sinnkd cos nkd sin nkd]

For the special case of half-wave spacing the result is:

Collinear short dipoles or slots yield a similar result:


(9 25)
6 iS U 1 [coswfa/ si l

Again the special case of half-wave spacing gives:

N
For both parallel and collinear short dipoles the mutual resistance is given by the
bracketed trig terms times a constant. Short elements are seldom used in arrays
because of their high reactance (susceptance) and narrow bandwidth; half-wave or
resonant elements are usually preferred. One might infer that directivity of half-
wave dipole arrays could be calculated using mutual resistance, and this proves to
be the case. The directivity will be derived for a uniform broadside array of half-
wave dipoles in any orientation. Directivity is the ratio of peak field power density
to input power:

G = "-^ (9.27)

For an array of half-wave dipoles the peak electric field Eo is:

where the dipole currents (excitation coefficients) are In. The numerator is then,
for uniform excitation (/„ = / 0 for all n):
IN \2
£ // JJ == UQN2^
num = 120 (( £ (9.29)

The input power is:

P = I Re (fnK) = I In I InRnm = 4 X E Rnm (9-30)


1 1 1 1 1
where the mutual resistance i$Rnm. Finally the directivity becomes:
18 Linear arrays
120N2 12(W
G = (9.31)

0 0-2 0^ 0-6 0-8 10 12


d/X
Fig. 9.14 Directivity of parallel dipole array

where Ro is self resistance, and Rn is mutual resistance between dipoles nd apart.


Directivity for parallel and for collinear arrays of half-wave dipoles has been calcu-
lated using compact efficient computer algorithms (Hansen71, Hansen and
Brunner73) with results shown in Figs. 9.14 and 9.15. It can be shown that the
Linear arrays 19

collinear case exhibits less grating-lobe directivity drop around d = A, but the peak
value is also lower. This is because the mutual coupling effect is greater with parallel
dipoles. The parallel array gives roughly 3 dB more directivity at the peak, due to
narrowing the beam in the transverse plane.

Fig. 9.15 Directivity of collinear dipole array

For fixed array spacing and number of elements, and for uniform amplitude of
excitation, the maximum directivity does not occur at broadside or endfire. This
case is included in Section 9.1.3 on superdirective arrays.
20 Linear arrays
An array with tapered excitation has a space factor that can be written as:

F(u) = I Anej2n™ (9.32)

Inserting this into eqn. 9.17J and integrating gives, for a tapered array of isotropic
elements:
' N

G
=

lm=l
Here it is assumed that the coefficients are real, but this restriction is removed
simply by using AnA*m,
When all ^4n = 1 this reduces to eqn. 9.20. If the spacing is d - A/2, the direc-
tivity becomes:

G = i '- (9.34)

Symmetric arrays allow eqn. 9.33 to be written with sums going to N/2 (oxN/2 +
1). In this case two sine terms are involved. As an example, the result for TV even is
shown:

G = (9.35)
N/2 N/2
Z L^n^m[ s i nc ( w + m
~ l)kd 4- sinc(« —rri)kd]
n = im = l

9.1.13 Tapered excitation

Dolph-Chebyshev arrays: A symmetrically tapered (amplitude) distribution


over the array or aperture is associated with a pattern having lower sidelobes than
those of the uniform (amplitude) array. Lowering the sidelobes broadens the beam-
width and lowers the excitation efficiency. The latter is the ratio of directivity to
the directivity for uniform excitation. Some improvement in both beamwidth and
efficiency is obtained by raising the farther-out sidelobes. Intuitively one might
expect equal-level sidelobes to be optimum for a given sidelobe level. A method of
accomplishing this for a half-wave spaced broadside array was invented by Dolph39,
who recognised that the Chebyshev polynomials were ideally suited: in the range
± 1 there are oscillations of unit amplitude, while outside this range the polynomial
becomes monotonically large (see Fig. 9.16). The Chebyshev polynomial can be
Linear arrays 21
expressed as:

( - 1 ) " cosh (n arc cosh |x|) x < -1

Tn(x) = cos(« arc COSJC) |x| < 1 (9.36)

cosh (n arc cosh x) x>1

Fig. 9.16 Chebyshev polynomial, T5{x)

However, a direct correspondence between this polynomial and the array poly-
nomial is not feasible, because the main beam must be symmetric and have zero
slope at its centre. The TV-element array has N— 1 zeros while the Nth order
Chebyshev polynomial has N zeros. Thus an N— 1 degree Chebyshev is used. Part
of the x > 1 region is mapped onto one side of the main beam while the oscillatory
portion of the Chebyshev is mapped out once onto the sidelobes on one side of the
main beam. The transformation from T^_l{x) to space factor F(\p), with i// =
kd sin 6, is x = XQCOS (I/ / /2). The voltage sidelobe ratio is given by:
SLR = (SLR>1) (9.37)
or inversely,
arc cosh (SLR)
x0 = cosh
N-l
and the space factor is simply
FOJO = 7WJOCOS0/2) (9.37a)
As B varies from —7r/2 to 7r/2, X varies from (x 0 cos nd/X) up to x0 at 6 - 0 and then
back to (x0co$ird/\). For half-wavelength spacing, the minimum value ofx is zero.
The maximum possible spacing is determined by the need to prevent* falling below
— 1. This gives the maximum element spacing for a broadside array as
22 Linear arrays

= arccos(-l/* 0 )
n
which approaches one-wavelength spacing for large arrays.
The excitation An of the nth element in the array may be obtained by choosing
the centre of the array as the phase reference and writing the space factor as
JV
Fty) = I Anew\j(2n-N- 1)^/2] (9.37c)
Since this Fourier series has a finite number of terms, its coefficients may be
obtained from the inverse relationship:

An = ™ I F{2itm/N) exp [-/(2n - JV - l)rtm/N] (937d)

or, substituting the Chebyshev polynomial representation:


1 N
An = TTI TN^(pc0cosmn/N) exp [-/(2« - J V - 1 W J V ] (9.37*)
iVm = l
Further simplifications may be made in view of the symmetric nature of the beam
and the excitation.
Formulas of Stegen146 are easily used for calculating the array amplitudes. These
number the elements from the array centre and are slightly different for odd and
even JV. Thus

An = | \sLR + 2 J rw_, t o c o s ^ j cos ^ j (0 < n <M) (9.38)


foroddA r =2M+l.and

An = 1\SLR + IfTs-i Lcos^) c o s ^ ^ l (Kn<M)


N N N
l \ ! J (939)
for even JV=2M
These formulas are awkward for large arrays; simplified approximate versions have
been developed by Van der Maas169 and Barbiere12. However, with current com-
puter capability, such approximations are unnecessary. Stegen's formulas can be
implemented directly. An extensive table of coefficients, directivity and beamwidth
has been prepared by Brown and Scharp24, giving N= 3(1)40 and SLR-
10(1)40 dB for d = A/2. However, there are significant errors near the upper values
of JV and SLRy presumably due to roundoff error. Users should compute their own
values using the formulas above, or those that follow.
Dolph's derivation and the formulas of Stegen are limited tod>X/2i because
only a small part of the oscillatory region of the polynomial is used with smaller
spacings. Riblet128 showed that this restriction could be removed, but only for JV
odd. For spacing below half wave, the space factor is formed by starting at a point
Linear arrays 23
near the end of the Chebyshev ± 1 region*, tracing the oscillatory region to the other
end, then retracing back to the start end and up the monotonic portion to form the
main beam half. Since the Afth order Chebyshev has M — 1 oscillations, which are
traced twice, and since the trace from 0 to 1 and back forms the centre sidelobe (in
between the trace out and back), the space factor always has an odd number of
sidelobes each side, or an even number of zeros. Hence only an odd number of ele-
ments can be formed into a Chebyshev array for d < X/2. The pattern is given by:
TM(a cos \p + 6 )
z0coskd
a = 1 — coskd ' b = ° (9.40)
coskd — 1
where, as before, \p = kd sin 0 and Af is (TV — l)/2. The value of z0 is different:
, [arc cosh (SLR)]
zo = cosh^ J (9.41)
SLR = TM(zQ)
Formulas have been developed by DuHamel,43 Bown,22>23 Salzer133 and Drane.40'41
Those of Drane will be used here as they are suitable for computer calculation of
superdirective arrays (see Section 9.1.3). The array amplitudes (numbered from
zero at the centre) are:

An=Z
4M \^M2-mTn{xn)[TM{axn + b)^(-\)nTM{b-axn)] (9.42)

where et = 1 for / = 0 and et — 2 for i > Q;xn = cos wnjM. The integers Mx and Af2
are, respectively, the integer parts of Af/2 and (M+ l)/2. This result is valid for
d < X/2. Small spacings (highly superdirective arrays) may require double precision
arithmetic due to the subtraction of terms. Many arrays are half-wave spaced; for
these the a and b of eqn. 9.40 reduce to:

^ , b = ^ (9.43)

For half-wave spacing the two approaches give identical results! In fact, due to the
properties of the Chebyshev polynomial, the two space factors, in precursor form,
are equal:

r , . . ( » f - « f ) . r , ( i & ~ » +•» + «»*-•) 0.44)


since, from eqns. 9.37 and 9.41, z$—2x%— 1. Since many computers have no
inverse hyperbolic functions, it is convenient to rewrite eqn. 9.41 as:
z 0 = i [SLR + > A ^ R 2 - 1 ] 1 / M + i [SLR -y/SLR2-l]1/M (9.45)

* The exact starting point depends on N and kd


24 Linear arrays

It was mentioned earlier that the Chebyshev distribution tended toward large peaks
at the array ends. To give an idea of the array lengths and sidelobe ratios for which
this occurs, the sidelobe ratio for the centre and end excitations equal is given
versus N, and also the value that ensures a monotonic distribution, in Table 9.1. For

Table 9.1 Chebyshev distributions

centre = end monotonic


N SLR, dB SLR, dB
10 16 22
20 20 33
30 22 —
40 24 -

a 20-element array, the centre equal-end situation occurs for SLR greater than
20 dB. Arrays larger than 28 elements do not allow a monotonic distribution at all
for SLR up to 40 dB. Fig. 9.17 shows array amplitudes for SLR = 20, 30 and
40 dB. Adjacent element amplitudes are connected by straight lines; but these
distributions are typical for Chebyshev arrays. In practice, Chebyshev distributions
are not often used, as the Taylor h distribution (see a later Section) offers more
flexibility.
SLR = 20dB

centre
array

Fig. 9.17 Chebyshev array distribution N = 19

Du Hamel43 extended the Chebyshev design principle to endfire arrays, but only
for d < X/2. In fact, to avoid a back lobe, spacing is customarily made < A/4. To do
this for any scan angle, \j/ is modified as usual to:
Linear arrays 25

\fr = kd(sme-sind0) (9.46)


where 60 is the scan (main beam) angle, and the interelement phase shift is kd sin 0O-
Coefficients a and b become:
3 -f z 0 -h 2\/2(z 0 4- 1) cosM
2 sin2 kd

b = (V^TT + f C ° S ^ (9.47)
2 sin kd
Drane42 has shown that a large scanned Chebyshev linear array has a directivity
independent of scan, for any spacings between half and full wave, for that scan
range for which only one main beam exists (no grating lobes). Directivity can be
calculated using the formulas of Section 9.1.1.2. Beamwidth is best calculated
directly from the space factor using a root finder such as Wegstein, with F(w3) =
F(0)/\/2. Approximate values may be obtained from calculations of Drane;42 Figs.
9.18 show directivity and beamwidth vs. SLR and L[\. It is interesting to note that
for each L/X there is an SLR giving maximum directivity. Smaller SLRs (higher
sidelobes) represent significant power contained in the equal-level sidelobes, while
larger SLRs yield lower efficiency due to the wider beam associated with lower
sidelobes.

Taylor n source: The continuous equivalent of the Dolph-Chebyshev array


space factor was developed by Taylor158 and was called the 'ideal' line source. This
distribution has a pattern with all sidelobes of equal level, so that intuitively it pro-
vides the narrowest beamwidth and highest aperture efficiency for a given sidelobe
level of any non-superdirective distribution. Equal-level sidelobes are provided by a
cosine function, while the main beam is provided by a hyperbolic cosine:
F(u) = co$hn^/A2 — u2, u<A
F(u) = co$ny/u2—A2, u>A
This is a one-parameter distribution because the parameters! controls all pattern
characteristics. Sidelobe ratio is simply the value at u = 0:

SLR = cosh?L4, A = - I n (SLR + ^SLR2- 1) (9.49)

Pattern zeros zn occur at zn = ± y/A2 + (n — | ) 2 . 3 dB beamwidth is:

u3 = -s/ln\SLR + y/SLR2 - 1) - ln2(SZ,JR/\/T+ yJSLR2/2 - 1) (9.50)

That the sidelobe envelope does not decay for large u is indicative of trouble, and
the aperture distribution corresponding to F(u) has a singularity (infinite value) at
each end; thus the appellation 'ideal'. Taylor invented a clever solution to this
dilemma. Close-in zeros should maintain their spacings to keep the close-in side-
lobes suitably low, thus keeping efficiency high. However, to reduce aperture
26 Linear arrays

energy storage Q, the sidelobe envelope for large u should taper as 1/w. To combine
these, the u scale is stretched slightly by a factor o slightly greater than unity, thus
maintaining close-in zero locations. At some point a zero can be made to fall at an

2OOO

1000

10 100 1000
sideiobe voltage ratio

10 100 1000
sidelobe voltage ratio

Fig. 9.18 Chebpshev array performance vs. SLR and LA. After Drane
a Directivity
b Beamwidth

integer; from this transition point the zeros occur at ±n. The pattern then has n
roughly equal sidelobes; beyond u — n the sidelobe envelope decays as 1/w.
Actually the first n lobes are not precisely equal; the transition region allows some
Linear arrays 27

decay of the sidelobes near n, and the envelope for u > ft is slightly different from
\ju there. Fig. 9.19 shows a typical pattern for SLR = 25 dB, and h = 5.
The zeros un become:
zn = ± (9.51)

Fig. 9.19 Taylor space factor


SLR = 25dB.fi= 5

The pattern function corresponding to these zeros can be written as an infinite


product on the zeros. But for both conceptual and computational reasons it is
advantageous to incorporate the infinite set of zeros into a trig function, leaving the
modification of n — 1 zeros in explicit form:

Fiu) = ^ ri (9.52)
nu n=i

F(n, A, n) sine TT(U + ri)


n=-(h-l))

where F(0,A,n)= 1. Thus the space factor can be viewed as a superposition of


n — \ beams each side of the central beam, where the beams are produced by uni-
form excitation and are spaced at unit intervals in u.
28 Linear arrays
The dilation parameter is:

Beamwidth is very closely given by oul3i where ul3 is the 'ideal' beamwidth. Table
9.2 shows the^l, w3 and o for various sidelobe ratios and values of n.

Table 9.2 Taylor modified line-source characteristics

SLR A u3 n=2 4 6 8 10
dB 0

20 0.9528 0.4465 1.1255 1.1027 1.0749 1.0582 1.0474


25 1.1366 0.4890 1.0870 1.0683 1.0546 1.0452
30 1.3200 0.5284 1.0693 1.0608 1.0505 1.0426
35 1.5032 0.5653 1.0523 1.0459 1.0397
40 1.6865 0.6000 1.0430 1.0407 1.0364
45 1.8697 0.6328 1.0350 1.0328
50 2.0530 0.6639 1.0289

The corresponding aperture distribution is best expressed in a Fourier series:


n-i
g(p) = 1 + 2 £F(w,,4,/0cosmrp (9.54)

where the aperture variable p is zero at the centre and ± 1 at the ends. The coef-
ficient is:
\(n-\V]2 "-1 n2
F(n9A,n) = , . / ; w - 1—TT n i - 4 - (9-55>
(n — 1 +n)l(n — 1 —ri)\ m = i z^
Tables of aperture distribution and coefflcents are given for SLR = 20(5)40 dB and
« = 3(1)10 by Hansen.69 More extensive tables have been calculated by Spell-
mire.142
It is necessary to discuss the appropriate range of values of n. Too large a value
of h will give a non-monotonic aperture distribution and may even produce a large
peak at the ends of the aperture. On the other hand, too small a value will not allow
the transition-zone zeros to behave properly. Fig. 9.20 shows distributions for the
largest values of n that allow a monotonic distribution: SLR = 25 dB, ft = 5; 30dB,
n = 7; and 40 dB, h — 11. A comparison of maximum efficiency and monotonic
distributions for SLR = 25dB is given in Fig. 9.21, where it can be seen that
n = 12 distribution is not only peaked at the end but has small oscillations on
either side of the dip. The difference in aperture efficiency, as will be shown next,
is only 1.5%.
The Taylor n distribution is widely used because it gives slightly better efficiency
and beamwidth than the Taylor one-parameter distribution, for the same sidelobe
level.
Linear arrays 29

Excitation efficiency (or directivity) of the n distribution is easily calculated.


Rather than integrating the space factor it is better to integrate the aperture distri-
bution since it is expressed as a Fourier series of orthogonal terms. The efflciencyis:
2
g(p)dp
(9.56)
\/ip)dp

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


aperture co-ordinate

Fig. 9.20 Taylor n aperture distributions

Note that this equation, which is often used, is valid only for distributions with
linear phase (Friis and Lewis58). It is not correct for difference-pattern distributions
for example. Using eqn. 9.54 for the distribution, the numerator is unity since the
cosine terms integrate to zero. In the denominator the product #2 before integration
gives unity plus two single series plus a double series. Again, each term of the single
series integrates to zero. The double-series integral is zero for n i= m, and gives \ for
n = m. Thus the excitation efficiency becomes:
1
T? = (9.57)

n-l
where the F(n9A9ri) coefficients are given by eqn. 9.55. Even large values of n
allow easy calculation of r\. There is, for eachSLR, a value of n that gives maximum
17. This is because, as it increases from a small value, the main beam energy decreases
30 Linear arrays
faster than the sidelobe energy increases. For larger n, however, the main beam
changes more slowly. Table 9.3 gives these peak efficiency values of n. These distri-
butions are not monotonic, and have significant edge excitation. Also shown in the
Table are the largest values of n that give monotonic distributions and the corre-
sponding efficiencies. It may be seen that the efficiency penalty is small.

max efficiency, n = 12

max monotonic, n r 5

02 0-4 0-6 08 1-0


aperture co-ordinate

Fig. 9.21 Taylor n aperture distributions for SLR = 25dB

Table 9.3 Taylor n distribution


max i? values monotonic n
SLR n r? n V
20 6 0.9667 3 0.9535
25 12 0.9252 5 0.9105
30 23 0.8787 7 0.8619
35 44 0.8326 9 0.8151
40 81 0.7899 11 0.7729

Taylor one-parameter source: As previously mentioned, few analytical tech-


niques are available for adjusting sidelobes of arrays. Usually a continuous distribu-
tion is used, with the array element values obtained by sampling the distribution.
Linear arrays 31

Other means for equating an array to a continuous distribution are discussed in


Section 9.1.2.2.
Following the principles of pattern control by zero placement, Taylor developed
a line source with adjustable sidelobe ratio and a l/« sidelobe envelope (low Q)
[Taylor;157 Hansen69]. Far-out zeros were left at the integers while close-in zeros
were shifted to reduce the close-in sidelobes. This was accomplished by setting
zeros equal to:
u = \Jn2+B2
where B is a real positive parameter. The canonical pattern function using products,
and with C a constant, is:

(9.58)
n i+- 2
n=l n
This can be more simply written as:

(9.59)

where the constant C has become sinh TTB/TTB. Thus this pattern is a modified
sinjc/x pattern; the single parameter B controls all characteristics: beamwidth,
efficiency, sidelobe level etc. A transition from the sine pattern to the hyperbolic
form occurs at u — B, on the side of the main beam. For smaller u the hyperbolic
form provides the central part of the main beam, with the peak value (u = 0) of
sinh-nB/irB. The sidelobe ratio is immediately that of the sinx/x times the beam
peak value. In dB,

SLR = 2 0 1 o g 5 ! 5 ^ + 13.26 dB (9.60)


TlB

Fig. 9.22 shows a typical pattern.


Table 9.4 gives for various sidelobe ratios values of the parameter B, beamwidth,
aperture efficiency r\ and edge taper. Since there is considerable interest in recent
years in very low sidelobe designs the table extends to SLR = 50 dB. The aperture
distribution is, of course, the inverse transform of eqn. 9.59:
g(p) = / o O r f f V P ^ (9.61)
where p is the distance from aperture centre (p — 1 at aperture end) and / 0 is the
zero-order modified Bessel function. At the centre the value g(0) = /O(7r2?) while at
the ends #(1) = 1; so the pedestal height is l//0(7rB) (see Fig. 9.23).
The aperture efficiency is given by:
32 Linear arrays

fF(0)]2 (9.62)
2
j_*[F(u)] du

Fig. 9.22 Taylor space-factor one-parameter, SLR = 25dB-

Due to the rapid sidelobe decay with u, the integral may be approximated with
infinite limits even for modest-length apertures. This allows the integration to be
reduced to a tabulated integral (Rothman;130 Abramowitz and Stegun1).
2sinh27r£ , ^x
(963)
)
where Jo is the integral of IQ from 0 to the argument. All characteristics of the one-
parameter distribution are thus readily available.
Linear arrays 33
Table 9.4 Taylor one-parameter line source characteristics

SLR B V Ws edge taper


dB rad dB
13.26 0 0.4429 1 0.443 0
15 0.3558 0.4615 0.993 0.457 2.5
20 0.7386 0.5119 0.933 0.478 9.2
25 1.0229 0.5580 0.863 0.481 15.3
30 1.2762 0.6002 0.801 0.481 21.1
35 1.5136 0.6391 0.751 0.480 26.8
40 1.7415 0.6752 0.709 0.479 32.4
45 1.9628 0.7091 0.674 0.478 37.9
50 2.1793 0.7411 0.645 0.478 43.3

E 0.4

0.2 0.4 0.6 0.8 1.0


aperture co-ordinate

Fig. 9.23 Taylor one-parameter aperture distributions

The half beamwidth u$ given in the Table is in terms of u. This is converted to


actual angle 6 by:

h M
-U> (9.64,

Table 9.5 gives values of normalised beamwidth for various SLR and L/X. For long
sources normalised beamwidth is independent of L/X, and for this case:

^ * 2a 3 (9.65)
34 Linear arrays

Table 9.5 Normalised beam width of Taylor one-parameter line source

SLR 00 L/X
dB
3 10 30 100
13.26 0.4429 0.8891 0.8862 0.8859 0.8859
15 0.4615 0.9267 0.9233 0.9230 0.9230
20 0.5119 1.0288 1.0242 1.0238 1.0238
25 0.5580 1.1225 1.1165 1.1160 1.1159
30 0.6002 1.2085 1.2011 1.2004 1.2004
35 0.6391 1.2880 1.2790 1.2782 1.2781
40 0.6753 1.3622 1.3515 1.3506 1.3505

1-5

13

12

1-1

10

09

0-8

0-7

0-6

0-5
10 15 20 25 30 35
sidelobe r a t i o . dB

Fig. 9.24 Normalised beamwidth vs. sidelobe ratio for linear one-parameter source

An SLR of 40 dB, for example, gives a beamwidth about 50% larger. Fig. 9.24
shows normalised beamwidth for sources 3 and 30 wavelengths long.
A comparison of Taylor h = 5 and Taylor one-parameter distributions for 25 dB
SLR is interesting. Figs. 9.19 and 9.22 show the space factors; the far-out sidelobes
of the one-parameter are roughly 5 dB lower. Excitation efficiency is, as expected,
higher for the h: 0.91 for n = 5 and 0.85 for one-parameter. Fig. 9.25 shows the
distributions. Surprisingly, one may note that these differ only in the 'tails', where
Linear arrays 35

the h has a much higher pedestal. Thus, the n distribution should be used where
efficiency (directivity) is crucial, and the one-parameter distribution should be used
where far-out sidelobes are important.

taylor n = 5

taylor one - parameter

center end

distance along source

Fig. 9.25 Aperture distributions for 25dB SLR

Beam Efficiency: For radiometer and some mapping applications, the fraction
of energy radiated in the main beam, null-to-null, is important. This quantity is
called the beam efficiency r?b. This Section determines how beam efficiency varies
with other design parameters such as sidelobe ratios, beamwidth, taper efficiency,
edge illumination etc. The vehicle for doing this is the one-parameter line source
developed by Taylor. The geometry is sketched in Fig. 9.26 where the line source
is along the z axis in spherical co-ordinates. The narrow pattern is rotationally sym-
metric about z, giving an integrand factor of cos0. Using the one-parameter space
factor F(u) developed previously, the beam efficiency is:
n
|F| 2 cos0tf0
(9.66)

Normally the evaluation of this would require numerical integration over all the
pattern sidelobes, an expensive and relatively inaccurate task. However, the taper
efficiency rjt, which is the ratio of directivity to directivity of the same source uni-
formly illuminated, can be accurately evaluated. These values can be used to evalu-
ate the denominator of the expression, leaving only a numerical integration over the
main beam. The latter is a simple and accurate task. The taper efficiency is:
36 Linear arrays

FIX sinhTri?
Vt = n/2
Fa = (9.67)
rn/2 nB
2L \F\2cosdd6
Jo
This gives:

Vb = )F\2 cos ode (9.68)

Fig. 9.26 Main beam integration

where the first null occurs at 0x. The dependence upon L/X can be removed for long
sources. ForX > X, dt < 1 and

lFldtt (9.69)
* FS J o >
Figure 9.27 and Table 9.6 give the values of beam efficiency vs. sidelobe ratio.
Other parameters are found readily from the value of B.

Numerical adjustment of sidelobes: Individual sidelobes, or groups of sidelobes,


can be raised or lowered by adjusting the relevant space-factor zeros. Of course if
sidelobes are lowered by moving zeros closer together, other sidelobes will raise
and/or the main beam will broaden, unless the aperture length is increased to add
zeros. This adjustment of zeros is conveniently done with the space factor written
as an infinite product on the zeros. If the pattern is asymmetric, the left and right
zeros must be expressed separately, of course. A perturbation method is then used
on the zeros to produce the desired pattern modifications. Elliott45"47 has, for
example, designed and tested a 19-element shunt slot array with a Taylor SLR =
20 dB, n = 6 space factor modified to have the first three sidelobes on one side
depressed to 30 dB. Fig. 9.28 shows the theoretical and experimental patterns.
Linear arrays 37

Table 9.6 Taylor one-parameter line source

SLR B 7]f Vb
dB
13.26 0 1 0.9028
15 0.3558 0.9931 0.9364
20 0.7386 0.9329 0.9820
25 1.0229 0.8626 0.9950
30 1.2762 0.8014 0.9986
35 1.5136 0.7509 0.9996
40 1.7415 0.7090 0.9999

0-9-

0-8
20 25 30 35
sidelobe ratio, dB

Fig. 9.27 Beam efficiency of Taylor one-parameter fine source

These excellent results reinforce the concept of designing space factors by proper
placement of zeros.

Bickmore-Spellmire two-parameter family: Sidelobe envelope taper can be con-


trolled in a two-parameter family of distributions developed by Bickmore and Spell-
mire.15 Both the Taylor 'ideal' and the Taylor modified sinm/mi space factors are
special cases. The two-parameter space-factor family is:
C2)
F(u) = (9.70)

cosine form of eqn. 9.48 results, while v = \ yields the one-parameter


38 Linear arrays

sin m/m of eqn. 9.59. The parameter C now controls initial sidelobe level, while v
controls envelope decay. Note that these two-parameters are independent, unlike
the obsolete multiple-parameter distributions mentioned earlier. The Bickmore-
Spellmire aperture distribution is:

-40
0° 30° 60° 90° 120° 150° 180°
angle from endfire

-10

-20

-30

-40
0° 30° 60° 90° 120° 150° 180°
angle from endfire

Fig. 9.28 Numerical adjustmen t of zeros {after Elliott and Johnson)


a Theoretical
b Experimental

g(p) = (9.71)
Linear arrays 39

Aperture end singularities occur for v < \, while v > \ gives envelope tapers more
rapid than 1/w, but commensurately less efficient.

Low-sidelobe designs: In the last decade low sidelobes (roughly —30 to — 60 dB)
have become of interest for several reasons: increasing spectrum congestion in
broadcast and in satellite transmissions, desirability of reducing radar intercept
probabilities, and reduction of radar clutter/jammer vulnerability. From the
previous Sections one may infer that a good low sidelobe distribution should be
heavily tapered in amplitude. Examination of the one-parameter space factor shows
very low pedestals for 40-60 dB SLR; from Taylor's observation that distributions
with pedestals are more efficient, the n space factor might appear more attractive.
This is indeed the case. To design a high-performance low-sidelobe space factor,
again the emphasis should be on the pattern zeros. Here the Taylor h space factor
is almost uniquely properly designed. Recapitulating, close-in zeros are adjusted to
obtain a few nearly equal sidelobes at the design sidelobe level, while farther-out
zeros are properly placed to give a \\u envelope. One may contrast this with a
Hamming distribution, which is a cosine of a doubled argument, on a pedestal. The
argument produces a zero contribution from the cosine at the centre of the half
aperture. This distribution is (Blackman and Tukey16):
g(p) = 0.54 + 0.46 cosTrp = a + b cosnp (9.72)
resulting in a centre value of unity and a pedestal (end value) of 0.08. Excitation
efficiency of the Hamming is:

= °- 7338
and the half-power beamwidth is u3= 0.6515. The corresponding space factor has
zeros at u = 2, 3, 4, . . . plus another zero at u = yja/(a — b) = 2.5981. The space
factor is:
= [(abWaUinnu
nau(u2- 1)
The close spacing of the first two zeros produces an irregular convex sidelobe enve-
lope where the highest sidelobe is the fourth at — 42.7 dB. First and second side-
lobes are —44.0dB and — 56 dB. Moving the salient zero to 2.5 to centre it does
lower the first and raise the second, but the overall SLR is now 1 dB worse.
A comparison of several low sidelobe distributions is illuminating. Take SLR =
42.7 dB for an example, since this is the Hamming value. Table 9.7 gives the first
ten pattern zeros for the Taylor one-parameter, the Hamming, and Taylor n = 6
and h = 10. All the Taylor distributions exhibit a relatively smooth progression; the
Hamming has the salient zero at u = 2.6. This is better seen in Table 9.8, where the
spacings between adjacent zeros are shown. Clearly, the Hamming is wasting zero
spacing on the first several sidelobes. In this Table can also be seen the effect of too
large an n for the SLR. n = 10 gives a slight oscillation in spacing. Finally Table 9.9
40 Linear arrays
Table 9.7 Zeros for SLR = 42.7 dB
n Taylor Taylor Hamming
one-parameter
fl = 6 « = 10
1 2.112 1.894 1.897 2
2 2.732 2.398 2.396 2.598
3 3.550 3.173 3.166 3
4 4.412 4.069 4.056 4
5 5.335 5.020 5.002 5
6 6.282 6 5.978 6
7 7.243 7 6.970 7
8 8.214 8 7.974 8
9 9.190 9 8.984 9
10 10.172 10 10 10

Table 9.8 Null spacings for SL R = 42.7 dB


n Taylor Taylor Hamming
one-parameter
n=6 n= 10
1 0.619 0.503 0.499 0.598
2 0.799 0.776 0.770 0.402
3 0.881 0.895 0.890 1
4 0.923 0.951 0.946 1
5 0.947 0.980 0.976 1
6 0.961 1.0 0.993 1
7 0.970 1.0 1.003 1
8 0.977 1.0 1.011 1
9 0.981 1.0 1.016 1
10 0.985 1.0 1.0 1

Table 9.9 Comparison of Distributions for SLR = 42.7 dB


Taylor Taylor Hamming
one-parameter
n =6 n= 10
"3 0.694 0.637 0.635 0.651
n 0.690 0.754 0.755 0.734
0.478 0.480 0.480 0.478
Linear arrays 41

gives beamwidths, efficiencies and products for these cases. As expected, the Taylor
h is the best. A choice of h can be made on the basis of aperture distribution shape
and pattern. The h pattern also is convex, but less so. Too large an n will also
increase the excitation tolerance requirements.

Average sidelobe level: In the last several years there have been new restrictions
on antenna sidelobes, from the standpoint of personnel hazards in the path of high-
power transmitting antennas, and to reduce interference between closely located
antennas. Thus both the sidelobe envelope and the average sidelobe level are
important antenna design parameters. It is desirable to be able to estimate the aver-
age sidelobe level for an antenna, and to know what sidelobe envelopes are reason-
able. This can be done as the average sidelobe level and envelope are approximately
related to the antenna directivity. This approximate relationship will be derived for
the Taylor one-parameter source. Since most antennas (linear) have a sidelobe
envelope that asymptotically decays as 1/w, the Taylor one-parameter source is an
excellent model. The average power sidelobe level is defined here as the integral
from the first null to TT/2:
L/K d
- _A_ [ "

where the first null is at u0 = \/l +B2. For long sources, L/\ > B, and
SLL Sl l n (976)
- ^ "^f
The actual sidelobe power is less because the sidelobes fit under the envelope.
Taking a sidelobe (power) shape of sin2?rw, independent of envelope, gives an
integral of 1/2. So the sidelobe power is 3 dB below the envelope. If the directivity
is G = 2T?Z,/A, with 77 the taper efficiency, the average sidelobe level may be written
in terms of the directivity:

SLLave
1 2i (9 77)
~ n^G W '
For a given initial SLR, B is fixed, and the right term can be calculated. This term is
shown in Fig. 9.29 vs. SLR and it can be noted that it varies from 0 to 2.7 dB for
SLR from 13 to 40 dB. The average SLR is then:

i? iIn
SLRave ^ G ^ + (SLR - 13.26 dB) - 10log ^ + 9.95 dB
2/
(9.78)
42 Linear arrays

Thus the average sidelobe ratio is 10 to 34 dB lower than the directivity (in minus
dB).
Edge pattern amplitude, at 6 = 90°, is also of interest. For large L/X the edge
field is simply X/nF0L, or in terms of directivity:
27?
(9.79)

-2-

dB

-3-

-5
10 20 30 z.0 50 60
SLR.dB

Fig. 9.29 One-parameter line source factor

In decibels this becomes


~(SLR- 13.26dB)+201ogr?-3.92dB. (9.80)

9.1.2 Pattern synthesis

9.1.2.1 Shaped beams: Shaped beams are usually used to cover an area; exam-
ples are the radar antenna with cosecant elevation pattern to compensate reduction
of range with elevation, or the satellite antenna pattern shaped to fit a given geo-
graphical area. Almost always it is the pattern amplitude that is important; the
phase is not prescribed. Assuming a phase distribution may require an inefficient
distribution, so only the amplitude (or power) pattern should be synthesised. A
series of orthogonal functions could be used, e.g. Fourier series, but many terms are
required as the functions do not fit the array physics gracefully (Jordan and
Balmain87). Weighting functions have been used to improve convergence for shaped
Linear arrays 43

beams (Bricout21) and for pencil beams (Jaeckle84). Chebyshev polynomials have
also been adopted for sector-pattern synthesis (Evans50). Where maximum flatness
is important the Butterworth-function method of Ksienski97 may be used. For
pencil-beam patterns the zero placement technique described earlier is preferable. A
better way of handling shaped beams analytically is by using orthogonal beams, fit-
ting a series of these beams to the desired shaped beam. The Woodward and Lawson
technique,180'181 although developed for continuous apertures, is equally useful for
arrays. Each beam is formed from a uniform-amplitude and linear-phase distribu-
tion, where the nth beam is given by sinNm/N sin m, where the array hasiV ele-
ments and u — (d/X) (sin0 — sin# n ). Here the array spacing is d and the beam peak
is at Bn. If the beams are spaced apart (in sin#) by nX/Nd, then all beams except the
nth will have zeros at the nth beam peak. This orthogonality allows the desired
pattern to be 'sampled' at approximately NX/2d points. For example, a half-wave
spaced array has zeros in sin 6 at 2n/N. A ten-element array can have ten beams
positioned, one at each null. Since the beams are orthogonal, the desired pattern at
each sampling point is exactly equal to the amplitude of the beam that peaks at
that angle, giving the amplitude immediately. The phase tilt for that beam corre-
sponds to the beam peak angle. And the total array distribution is simply the sum
of these for each constituent beam. If the desired pattern is F(w), and if the beam
peaks are at un and of amplitude F(un), the actual pattern is simply:

The amount of ripple will be controlled by the element spacing. Fig. 930a from
Woodward and Lawson181 is an example of a cosecant synthesis using 21 beams
(a 21-element array at half-wave spacing) where one zero is placed at Q = 0. Smaller
ripples will be obtained if the zeros are shifted to allow the cosecant drop-off to
cross the 0 = 0 axis; (see Fig. 930b). The element-spacing tradeoff involves pattern
ripple; larger spacing means fewer sampling points, but, more important, spacing
larger than half-wave will allow grating lobes (see Section 9.1.1) to appear if the
desired pattern must be sampled at wide angles. Spacing smaller than half-wave gets
into the realm of superdirectivity (see Section 9.1.3) and is generally undesirable.
Half-wave spacing, then, is a good choice for cosecant-type patterns.
The Woodward-Lawson synthesis used constituent beams that were derived
from a uniform distribution, giving for long arrays a —13.3 dB sidelobe level. It is
possible to start with the zeros of a low sidelobe distribution, then to move the
zeros off the unit circle (in or out) to shape the beam.
However, the orthogonal beam spacing here is larger; so the cost of achieving
lower sidelobes (and higher directivity) on the opposite side of the cosecant
pattern, for example, is larger pattern ripple or a larger aperture.
Since analytical synthesis does not control pattern ripple in the desired region,
and does not allow sidelobes outside the desired region to be independently speci-
fied, numerical methods have been developed which allow optimisation of both
features. In fact, it is not necessary that the pattern exhibit zeros (nulls). The array
44 Linear arrays

180°
amplitude phase

(a)

csc6

-10 -0-5

180°
amplitude 90°

-5 -3 -1
y/x 90 (

-180
- 5 - 3 - 1 0 1 3 5 phase
y/X

1-0J
(b)

csc8

-1-0 -0-5 0 05 10
sinG

Fig. 9.30 Cosecant synthesis (after Woodward and Lawson)


a Zero at broadside
b Improved
Linear arrays 45
polynomial zeros may be located off the unit circle (see Section 9.1.1). Intuitively
one might expect, for example, the optimum cosecant pattern to have nulls in the
low part of the pattern just adequate to give the prescribed sidelobe level. When the
entire pattern can be specified, for example in broadcasting antennas, a least-
squares error function is then minimised by an iterative scheme: Newton-Raphson
(see Section 9.2.1) or gradient (Perini and Idselis119). Of gradient algorithms, the
Fletcher-Powell53 and Fletcher-Reeves54 are widely available; significant improve-
ments have been made by Fletcher55 and Gill and Murray.63 Least-squares synthesis
tends to produce unequal ripples; an approach to a true minimax synthesis results
from a least Pth synthesis (Temes and Zai;159 Bandler and Charalambous11). These
techniques are not useful when sidelobe regions are involved because the number
and position of sidelobes is usually not important; only the sidelobe envelope is
important. Thus, the individual sidelobes cannot be effectively specified.
Another numerical synthesis technique uses iterative sampling (Stutzman150'1S1);
this method can be applied to both shaped beam and sidelobe regions. It is a
modification of the Woodward and Lawson method. First a pattern close to the
desired pattern is selected, or synthesised by analytical means. Next, a set of correc-
tion patterns is added to the starting pattern. A convenient correction pattern is the
uniform array pattern sinNml(Nsinm); a number of sample points on the pattern
are selected, and one correction pattern is applied at each sample point. The ampli-
tude of each correction pattern is proportional to the difference between the pat-
tern and the desired pattern at that sample point. Only if the samples are uncorrela-
ted will the resulting pattern match the desired pattern. However, if the sample
points are not close the correlation will be weak. A good spacing to use is to put the
peak of one sampling pattern at a null of the adjacent sampling pattern, just as in
Woodward-Lawson with sine patterns. Following this first iteration, a new set of
sample points is used; a good choice is each point midway between points of the
previous iteration. Fig. 9.31 shows a 5X aperture and a desired cosecant pattern,
with a Woodward-Lawson synthesis and two iterations. The directivities of the
latter are higher because the pattern does not overshoot and because the sidelobes
are lower. When sidelobe regions are involved a sample point should be at each side-
lobe peak of the starting pattern. As the iterations progress, the patterns should be
examined to see if the sidelobe sampling points need to be moved. Since the mathe-
matics and equations are simple, this iterative sampling method is fast and effective,
and can be used for adjusting individual sidelobe heights.

9.1.2.2 Sidelobe envelope shaping: Narrow-beam line-source patterns may have


sidelobe envelopes shaped or adjusted by the iterative sampling method just des-
cribed. A more direct method uses a Newton-Raphson type of iteration (Elliott45).
The starting pattern is first selected to be close to the desired pattern. It is con-
venient to write this pattern as a product on the zeros. Then each zero is shifted a
small amount and the resulting pattern written. From this sidelobe peak heights can
be calculated, and subtracted from the respective desired sidelobe heights. The
resulting set of equations is nonlinear and can be solved by Newton-Raphson
46 Linear arrays

— Fd(u)

— Fd(u)

(b)

— Fd(u)

(c)

Fig. 9.31 Iterative sampling synthesis [courtesy Stutzman)


a Woodward— Lawson
b First iteration
c Second iteration
Linear arrays 47
iteration. This method is effective because moving nulls apart raises the sidelobes.
Fig. 9.32a shows a space factor designed for three ~40dB sidelobes adjacent to the
main beam, then four sidelobes at — 20 dB, followed by the normal 1/w taper. In
Fig. 932b is shown the distribution required. The phase is constant while the
amplitude has a large dip near the ends. Although this example is symmetric, the
method applies equally well to asymmetric patterns.

0i
A
-10
l\
[
-20
db
AHA

1 I
-30
flAn/w
if
-40

Rfi
-U-12-10-8 -6 -L -2 0 2 4 6 8 10 12 U
z

1-0

0-8
T3

I 0-6

4, 0-4
>

? 0-2

-TT TT
-n 2 p=0 2
TT

20n
argf (p) phase

20

Fig. 9.32 Sidelobe envelope shaping, {courtesy: Elliott)


a Space factor
b Aperture distribution

When large arrays are used a line-source distribution can be sampled to obtain
element excitation values. For small arrays, the sidelobe envelope shaping must be
done directly on the array polynomial zeros, TV— 1 of them for an TV element array.
48 Linear arrays
These are complex zeros in the variable w = exp (j2nu). The complex coefficients
of this polynomial are just the element excitations. Newton-Raphson iteration is
again used to adjust each sidelobe to the desired peak value by perturbing the com-
plex zeros. However, only the sidelobe magnitudes can be specified, so the
complex-array polynomial must be normalised so that it is real for N odd (for N
even the complex polynomial is imaginary). Using this method questions of sampl-
ing accuracy are obviated, with excellent agreement-with actual arrays (Elliott46).
Another application of sidelobe envelope shaping through zero placement is the
production of a broad null, a few sidelobes wide. To accomplish this, the sidelobes
around the null portion are progressively lowered to simulate a broad null
[Tseng163]. Fig. 9.33 shows the modified pattern along with the starting 40 dB
Taylor space factor.

9.1.2.3 Constrained synthesis

Maximisation of G or SNR: Often it is advantageous to synthesise a high-


directivity (narrow beam) array, subject to a constraint. For example, a maximum
sidelobe envelope might be imposed; or a maximum Q or tolerance variance. The
latter are important when array spacing is less than a half wavelength. Synthesis of
maximum directivity without any constraints is discussed in Section 9.1.3, Super-
directive Arrays. That formulation will be seen to be a special case of the general
formulation to be developed here. Constrained synthesis is closely related to super-
directivity, as the constraints are often imposed to minimise the adverse effects of
superdirectivity. Many references are given by Bloch, Medhurst and Pool;17'18 they
also expressed directivity in terms of array mutual impedances. The basic frame-
work of constrained optimisation was developed by Gilbert and Morgan62 and
Uzsoky and Solymar,168 and was extended by Lo et al.10s A review paper is by
Cheng.27 Use of the bracket notation introduced by Dirac (Friedman57) allows the
formulation to be simplified. Let <A\ be a row vector and \A) be a column vector.
The scalar product then becomes:

UW> - I \An\2 (9.82)


where * indicates the conjugate. Call the row vector of complex array excitations
/ and the column vector of path-length phases F:

exp(—jkri sin0)
F = (9.83)

where rn is the distance from the reference point to the /tth element. For a uni-
formly spaced array, rn~(ji— \)d. Now define matrices^ and B, where A is:
\A\ = \F*)(F\ (9.84)
and the elements of B are:
Linear arrays 49

'o.oo 8.00 16.00 24.00 32.00 40.00


angle deg

0.00 10.00 20.00 30.00 40.00 50.00 60.00

Fig. 9.33 Broadsidelobe null (solid); 40dB Taylor (dashed) {courtesy: Tseng)
a Space factor
b Distribution
50 Linear arrays

Bnm = ^\fi(P> *)exp [-fkd(n-m)sinO]da (9.85)

The pattern of the zth element isft. Isotropic elements and a uniformly spaced array
allow simplifications of A and B. The elements then become:
(9.86)
Bnm = sine (n — m) kd
Now the directivity Q and sensitivity can be written in abbreviated form:

The tolerance sensitivity S is defined as the ratio of variance of peak field strength
produced by errors of variance o\ (Uzsoky and Solymar168):
(AE)2 oj(J\J*)

" 2 2 2
Oj OJ OJ
It may be noticed that Q = SG. The directivity is a ratio of two Hermitian quad-
ratic forms,^ with B positive definite and A at least positive semi-definite. Thus all
the eigenvalues of the associated equation
\A\J) = Gmax\B\J) (9.89)
are zero or positive real. Since A is a single-term dyad there is one non-zero eigen-
value. The eigenvector is found to be:
\Jmax) = \B\~lF) (9.90)
The corresponding maximum directivity is given by:
Gmax = (FlBr'F*) (9.91)
These compact results are the same as those developed in Section 9.1.3.3.

Maximisation of G or SNR with constraint: In many applications it is import-


ant to maximise G/T, directivity/system noise temperature. This is equivalent to
maximising signal/noise ratio SNR. To do this the element pattern in the integral
for Bnm (eqn. 9.85) is multiplied by the antenna noise temperature T(6, 0). Then
the excitation vector that maximises SNR or G/T is that of eqn. 9.90.
Another ratio that can be directly minimised is beam efficiency: the fraction of
power contained within the main beam, null-to-null. This is:

U W r >
(9.92)

' See Gantmacher61 for a general discussion of operators in Hilbert space.


Linear arrays 51
where B is as before, and A is the form of B with the limits of integration covering
only the main beam.
To avoid the triple problems of superdirectivity (see Section 9.1.3) maximisation
of G or SNR subject to a constraint on Q or S is advantageous. Or it may be desired
to maximise G or SNR subject to a maximum sidelobe envelope. Since all of these
constraint problems are basically the same, the process will be described for maxi-
misation of directivity subject to a general constraint of the form:

This allows Q, tolerance sensitivity and sidelobe envelope to be included as special


cases. The last will be discussed later. Following the development of Uzsoky and
Solymar,168 Lo et a/.,105 and Mautz and Harrington,109 a Lagrangian multiplier A is
used to form:

where ^ o is the specified value of the constraint. The variation of this expression
is now taken with regard to |/> and A, and set to zero. The result is:
|/> = q\KYl (9.95)
The matrix AT is:
|tf| = \B\+p$f\D\-p\C\ (9.96)
and the constants p and q are given by:
p = KU\B\J*)2U\A\J*rl{J\D\J*Tl
q = (J\B\J*)(J\A\J*Tl(F\J) (9.97)
with \A\ = \F*)(F\ as before. When these are substituted into the ^ equation the
result is the characteristic equation in p:
<F\K\'l( %f*\D\ - \C\)\KYlF*) = 0 (9.98)
The equation indicates that (F\ is orthogonal to the vector:
\K\~\ Mi\D\ - \Q)\KrlF*) (9.99)
A complete set of functions can be constructed with F as its first element. If
F= [/i,/ 3 ,...,/ w ] (9-100)
105
then the set is simply (Lo et al. ):
F, = [/I,/„ • • • ,fN]

= [-7,7,0,...,0
h h
52 Linear arrays
1

(9.101)

Now the vector (eqn. 9.99) can be expanded in the functions F2 through FN:
N
l
\K\-\ ^0\D\-\C\)\K\~ Ft) = £ ht\Ft) (9.102)
i=2

This can be written as a matrix times an a vector, and the determinant of the
matrix set to zero. The resulting polynomial in p has one root which can be found
by a rooter such as Newton or Wegstein. A faster procedure, since rooters only
work properly within a narrow range of starting values, is to separate the matrix
real and imaginary parts (Winkler and Schwartz 178 ) into a double-size matrix where
eqn. 9.102 is written as

\W\H) = 0, \H) = [~l,h2,h*...,hN] (9.103)

Wllr-Wm h \r
Wl2i~Wl2r hH
W21r-W21i • h2r = 0
W2U-W21r hti

This can be rearranged and rewritten as:


p2l^il+pl^2l+ l^l^*) = 0 (9.104)
The matrices U are W x 2N, and Ut and U2 have the first two columns zero. Solu-
tions of p are now real eigenvalues of the equation:

\V\X> = -IX]
P
where V is a 4N x 4N real matrix given by:

0 /
W\ = (9.105)

/ is the identity matrix as usual. Numerically it is much easier and faster to deter-
mine the eigenvalues of a 4N x 4/V real matrix than to find the roots of aniV-degree
polynomial. The set of values of p (the eigenvalues) must now be tested to find
which one maximises G subject to <3f= J%^. The process of determining the ps
and finding the correct one is lengthy but straightforward. Since most arrays to be
Linear arrays 53
optimised are not large, the matrix inversion and eigensolution processes are not
burdensome.
The generalised constraint (eqn. 9.94) can utilise Q, the supergain ratio, toler-
ance sensitivity S, sidelobe envelope etc. The sidelobe envelope constraint is
handled by defining a desired sidelobe-level vector (S\ which relates the amplitude
of each sidelobe to the beam peak:
(J\Fp(SL\ = [</|Ff>, U\F$\ . . . , </|F£>] (9.106)
To find the angles corresponding to the sidelobe peaks, values from the closest
canonical distribution are used. After the optimisation it may be necessary to re-
determine sidelobe peak angles and to re-run the optimisation. The general con-
straint matrix is now used with ^ = 0 .
Minimax optimisation has also been used (Mucci et a/.112), but these algorithms
are usually extremely slow when the number of variables is more than a few. Also
the least Pth gradient method is nearly equivalent and much faster as previously
mentioned.

Phase-only optimization: Since it is relatively easy to feed arrays uniformly, it


is important to be able to optimise performance by adjustment of only the phases.
Using the same performance indices that are quadratic Hermitian ratios, the
gradients can be written, where directivity is used as an example (Voges and
Butler170):
2Im<^^l/*>
d(t>i U\B\D
Now gradient methods referred to in Section 9.1.2.1 can be used directly with the
formulas above. Another method, which is faster, is an iterative gradient method,
Newton-Raphson (see Section 9.2.1). The directivity then becomes:
U\TAT*\J*> = (JC\AC\J*)
U\TBT*\J*) UC\BC\J*)
All of the previously developed optimisations can now be applied. However, it must
be remembered that there are only N— M degrees of freedom left to use in maxi-
mising directivity.
Synthesis of power patterns is straightforward, but the process of handling a
product of two sums instead of a single sum is tedious and lengthy.
A word on apertures is in order. Of course, a continuous aperture may be sam-
pled or discretised, which converts it to an array problem. The first solution of
maximum directivity for a line source subject to a Q constraint was obtained by
Rhodes,124*126 who recognised that a doubly orthogonal set of functions would
allow a direct solution. Orthogonality over a finite range allows power over real
space to be obtained, while orthogonality over an infinite range gives stored energy
over real and invisible space. Such functions are the prolate spheroidal wave func-
tions which are also square integrable. A Lagrangian multiplier is used for the
54 Linear arrays
maximisation of directivity. More general constrained synthesis is possible, but the
matrix manipulation methods applicable to Hermitian forms are not available.
Typically gradient methods are used (Evans and Fortmann51).
Each gradient is set to zero and a gradient matrix is formed by differentiating
each with respect to each phase. Inverse of this matrix times a start vector yields a
new vector which is iterated.
Optimisation with fixed excitation and variable array spacing is handled simi-
larly except that the gradient expression is slightly different.

Miscellaneous: Mutual coupling effects may be included by writing the direc-


tivity or other performance index in terms of mutual resistance as in Section
9.1.1.2. Loss may also be included in the formulation so that gain, instead of direc-
tivity, may be maximised.
Optimisation of directivity or SNR with nulls in prescribed directions can also be
accomplished. A null is equivalent to the excitation vector </| being orthogonal to
the vector (F*\, where the null is in the z'th direction. Thus for M null directions,
where M<7V, </| is orthogonal to <F*|. Then </| must lie in the remainder of the
complete orthonormal space that includes the M constraint vectors. An ortho-
normalisation procedure such as Gram-Schmidt (Hildebrand78) can be used to con-
struct this set. If the transformation matrix is \T\, the excitation vector is trans-
formed to the constrained vector:
\T\J) = \JC) (9.109)

9.1.3 Superdirective arrays


A useful operational definition of array superdirectivity (formerly called supergain)
is directivity* higher than that obtained with the same array configuration uni-
formly excited (constant amplitude and linear phase). Excessive array superdirec-
tivity inflicts major problems in low radiation resistance (hence low efficiency),
sensitive excitation and position tolerances, and narrow bandwidth. Superdirec-
tivity applies to arrays in two ways: to the array (of isotropic elements) and to the
element itself. Superdirective elements are those with dimensions much smaller
than X/2. They also suffer low radiation resistance, sensitive tolerances and narrow
bandwidth. The directivity is close to 1.5; thus using the uniform line-source direc-
tivity for comparison (G = 2L/X), there is a directivity increase from that value
(since L<\) to 1.5. And so electrically small elements are analogous to super-
directive arrays in all respects. Both are important in the restrictions imposed and
each will be discussed after some historical notes.

9.1.3.1 History of superdirectivity: Probably the earliest work^ on the possi-


bility of superdirectivity was by Oseen.117 A limited endflre superdirectivity using a

Directivity is the ratio of peak field intensity at any far field radius from the antenna to the
integral of field intensity over a sphere of that radius.
f See Bloch et al.17 for a list of early references.
Linear arrays 55
74
monotonic phase function was accomplished by Hansen and Woodyard. Another
early contributor was Franz.56 Schelkunoff,135 in a classic paper on linear arrays,
discussed, among other topics, array spacings less than A/2, showing how equal
spacing of the array polynomial zeros over that portion of the unit circle repre-
sented by the spacing gives superdirectivity. The field received wide attention when
La Paz and Miller100 purported to show that a given aperture would allow a maxi-
mum directivity, and when Bouwkamp and De Bruijn19 showed that they had made
an error and that there was no limit on theoretical directivity. Thus the important
theorem: a fixed aperture size can achieve (in theory) any desired directivity value.
This theorem is now widely recognised, but the practical implications are less well
known. Bloch, Medhurst and Poole17 say that the theorem has been rediscovered
several times; the practical limitations of superdirectivity occur as a surprise to sys-
tems engineers and others year after year! In 1946, a burst of wartime research
reporting occurred. Reid123 generalised the Hansen-Woodyard endfire superdirec-
tivity to include an element pattern. Uzkov167 derived the endfire directivity as
d -> 0. And Dolph39 invented the widely used Dolph-Chebyshev array distribution
wherein the equal-level oscillations of a Chebyshev polynomial are used to produce
an array pattern with equal-level sidelobes (see Section 9.1.1). To follow this last
development, Riblet127 developed Dolph-Chebyshev arrays for spacing below A/2,
i.e. superdirective. DuHamel43 and Stegen146 developed complementary advances
in the computation of Dolph-Chebyshev coefficients and directivity. Chu30
developed fundamental limitations on electrically small antennas, deriving Q and
directivity from a consideration of spherical modes. Maximum directivity for an
array with fixed spacings was derived, for acoustic arrays, by Pritchard.120

9.1.3.2 Element superdirectivity: With the miniaturisation of components


endemic in almost all parts of electronics today, it is important to recognise the
limits upon size reduction of antenna elements. These are related to the basic fact
that the element's purpose is to couple to a free-space wave, and the free-space
wavelength has not yet been miniaturised! A basic approach was taken by Chu30
and subsequently by Harrington.75 Since any radiating field can be written as a sum
of spherical modes, the antenna, of whatever type it happens to be, is enclosed in a
sphere. The radiated power can be calculated from the propagating modes within
the sphere. All modes contribute to the reactive power. When the sphere is
sufficiently large to support several propagating modes, this approach is of little
value as the modal coefficients are difficult to calculate. With only one propagating
mode, the radiated power arises primarily from that mode, analogous to the unit-
cell analysis developed by Oliner and Malech115'116 for an electronic scanning
(phased) array antenna. The utility of the Chu work becomes apparent when the
sphere is too small to allow a propagating mode; all modes are then evanescent
(below cutoff) and the Q becomes large, as the evanescent modes contribute little
real power. Note that, unlike a closed waveguide, there is a real part of each evan-
escent mode. Each mode has a Qn based on the ratio of stored energy to radiated
energy and the Qn rises rapidly when ka drops below the mode number. The overall
56 Linear arrays
Q of an idealised optimum directivity antenna is given by:

I (2« + l)Qn
Q (9U0)
* "'W + 2)
For electrically small entannas or arrays, we are concerned primarily with one
mode, for which the overall Q is half the mode Qn (equal power in TE and TM
modes)* The single mode antenna is given by:

which is plotted in Fig. 9.34. Note that for k < 1, the Q varies inversely as the cube
of ka or array length in wavelengths. This result corresponds to a single antenna
element. Exciting the element equally with electric and magnetic mode halves the Q
for a given ka. Also shown are curves with efficiencies below unity. Since the loss is
just a resistance in series with the radiation resistance, its effect on Q is apparent.
From these curves then it can be seen that a single mode element (as most are)
cannot be much below half wave in size for an octave bandwidth (Q = 1.414). And
small elements will have a narrow bandwidth especially since most elements do not
use the fundamental spherical volume efficiently. Size reduction of elements is thus
usually impractical.
Bandwidth is derived from Q by assuming that the antenna is a resonant circuit
with fixed values. Then bandwidth is:

gyy = fupper ~~ flower _ J_ (9.112)


/centre Q
This allows a 3 dB power decrease at band edges. For Q > 1 this relationship is
accurate, but for Q = 1 or 2 a simple resonant circuit is not a good representation
of any antenna.

9.133 Array superdirectivity

Maximum directivity for uniform amplitude, fixed spacing: In 1938 Hansen and
Woodyard74 developed an endfire line source with modest superdirectivity. This is
of interest because the distribution can be sampled to get array excitations and
because the amplitude is constant, a feature that is attractive for arrays. They
observed that, if the free-space phase progression along the array were increased,
the space-factor power integral decreased faster than the peak value; thus the direc-
tivity increases up to a point. The endfire pattern is:

F(d) = sine ~ ( i t c o s 0 - 0 ) (9.113)

where 0 is now measured from endfire and j3 is the wave number over the aperture.
Inverse directivity is proportional to:
Linear arrays 57
1
— oc *- s f sine2 -(fccos0—0) sinBdd
G 2
sine -<k-8)\
-(* ° L J

1 l^.o^. cos0-1
(9.114)
2

0.1 0.3 1.3 1.5

Fig. 9.34 Chu—Harrington Fundamental limitations for single-mode antenna, various


efficiencies
58 Linear arrays

Here Si is the sine integral, and <f> = L(k — p)9 the additional phase along the aper-
ture (in addition to the progressive endfire phase). Maximum directivity was deter-
mined to occur for $ — 2.922 radians. In many books it is carelessly stated that n
extra radians of phase are needed, but there is no physical reason for this; a better
approximation to 2.92 is 3. Directivity increase over normal endfire is 2.56 dB, and
the sidelobe ratio is 9.92 dB. This distribution is suitable for long arrays; for short
arrays a computer optimisation of phase is recommended.
n W2 w 3w/2 k o d=2 7r

2-25

20
-TT/2

1-75 20

Fig. 9.35 Directivity of Uniform array [courtesy: H. Bach and J. E. Hansen)

b A/=3
The Hansen-Woodyard distribution is endfire. In general, the maximum direc-
tivity does not occur there. The most general solution for uniform amplitude would
allow any element phases needed to maximise directivity. Such a solution could be
formally realised for a given number of elements and spacing, but the equations
would require a numerical solution. A slightly simpler problem was worked by Bach
and Hansen;9 they started with a uniform-amplitude array that was phased to pro-
duce a main beam at 6Q. The interelement phase factor is 6:
5 = kd sin 60 (9.115)
and the directivity (see Section 9.1.1.3) is:
Linear arrays 59

s-y N
(9.116)
1 + 7; Z (W — n) sine (wfaO cos (nb)

Calculations were made for 2-, 3-, 4- and 10-element arrays for all beam angles, and
for spacings up to X. Fig. 9.35 shows their results.
Fig. 9.3Sd is striking in that high directivity occurs along a line roughly for

37T/2 k O d=2rf

-it/2 - '

ir\l •T 3fr/2
]
v*—"^^ y3
1 '
^7
9
-Tt/2-
12
8 ^ ^ S8 10

8=-Tt I

(d)

Fig. 9.35 Directivity of Uniform array (courtesy: H. Bach and J. E. Hansen)


cN = 4
dN= 10

A;d + 5 = 0, or 0O = ~TT/2, with peak directivity near endfire at X spacing. Along the
line roughly for kd = 5 + 2TT directivity is changing rapidly, perhaps due to appear-
ance of another lobe. Directivity values are shown at 0O> but in some cases a 'side-
lobe' may have higher amplitude. Thus even for uniform amplitude an array is
complex. Next variable amplitude excitation is allowed.
60 Linear arrays

Maximum directivity for fixed spacing: For an array with a given number N of
elements and a fixed spacing, it is clear that there is a maximum directivity. The
array coefficients that produce this maximum are found by the Lagrange multiplier
method (Courant31). Since the set of coefficients can be scaled by any common
factor, it is convenient to make the sum unity. Then for an array of an even number
of elements:
N/2
Y.An = 1 (9.H7)
n=l
This allows the inverse directivity of a broadside array (which will be minimised) to
be expressed as:
i N/2 N/2
~ = 2 I I AnAm[fnm+fn.m] (9.118)
tr n-im = l
where fn = sine (nkd). An odd number requires a change of limits and introduction
of the Neumann number.
Applying the variational method gives a set of equations, where the Lagrangian
multiplier is/*:
N/2
I i4B [ / „ • „ - ! + / » - m ] + P = 0, m=l,2,...,Nl2 (9.119)

The set of equations hasjV/2 + 1 unknowns andiV/2 equations; again the sum equa-
tion completes the set. The multiplier can be eliminated by substituting from the
m = 1 equation.
N/2
M = - lAn[fn+fn-i] (9.120)
n=l
This gives a set of N/2 simultaneous equations in the array coefficients. These were
solved by hand for iV = 3, 5, 7 by Pritchard,120 and can readily be solved by com-
puter for any reasonable N. Writing the set of equations:
N/2
I An = 1
n=l
N/2
YAn[fn + m-l+fn~m-fn~l-fn] = 0, m = 2, 3, . . . ,M (9.121)
n=l

The RHS vector is (1, 0 , . . . ) and the matrix of unknown coefficients is:

"J"/n-3~~/n-l ~~ fnl

-1 +fn-M~fn-l ~ fnl (9.122)


Linear arrays 61
Here each row is filled out by letting n = 1 , 2 , . . . ,N/2. Although matrix inversion
can be used to solve for the array coefficient vector, it is much faster to use an
equation solver such as the Crout reduction (Westlake174).
Calculations of maximum broadside directivity were made with the results
shown in Fig. 936. Also shown is the directivity for uniform excitation. Above
d/\= 0.5, the two are very close. Also some minor oscillations in the directivity
curves have been smoothed out, as they are not important here. The coalescing of
pairs of curves at zero spacing will be commented on in the next Section.

03 CU 05 06

Fig. 9.36 Maximum directivity for fixed spacing


Uniform amplitude

For endflre beams, in the directivity expression (eqn. 9.118)/ n is replaced byf2n
and this carries over to the simultaneous eqns. 9.119. The element coefficients will
now be complex, due to the Hansen-Woodyard endflre aperture. For progressive
phasing (phasing matches space-wave phasing), the directivity is N for spacing of
both X/4 and X/2. However, at X/4 spacing, the optimum excitation in amplitude
and phase yields a directivity of iV2.

Maximum directivity as spacing goes to zero: Uzkov167 transformed the maxi-


mum directivity limit (as spacing goes to zero) for an array into a sum of Legendre
polynomials. His result for an array of N elements is:
62 Linear arrays

(9.123)

where 6 is the angle of the main beam from broadside. Tai154 independently
developed this for broadside arrays. Here cos0 = 1. Since the Legendre polynomial
at argument 1 can be written as a product of factors, the result for maximum direc-
tivity is simply
I . 3. 5 . . . E N+ 1
where E = 2 x AINT -1 (9.124)

Here E is the number of odd elements or the number of even elements minus 1.
Thus the interesting result that 3-and 4-element arrays have the same limiting value,
5- and 6-element arrays have the same value, etc. After reflection this is not surpris-
ing as both 3- and 4-element arrays have two variables, both 5- and 6-elements have
three variables, etc. This maximum directivity is plotted in Fig. 9.37. The circles
show the corresponding limiting directivity for Chebyshev arrays with lOdB SLR
and the squares show for the 20 dB SLR. The Chebyshev directivity is less, as the
maximum directivity pattern does not have equal-level sidelobes, and even in the
case where there is only one sidelobe (backlobe), the Chebyshev result would be
equal to the maximum value only if the sidelobe ratio were properly chosen. See
the next Section for a discussion of Chebyshev arrays.

9 10 11 12 13

Fig. 9.37 Maximum directivity in limit of zero-spacing broadside


® Chebyshev, SLR = 10dB
B Chebyshev, SLR = 20dB

For endfire arrays, the maximum directivity in the limit of zero spacing is:

G -> X On ^ l)[Pn(l)]2 = Z 2n 4- 1 = N2 (9.125)


n=0 n=0
Linear arrays 63

Uzkov also shows that the maximum endfire directivity at X/2 spacing is N. For
X/4 spacing, the maximum directivity isN2.

9.13A Dolph-Chebyshev arrays: There are several ways of synthesising an


array pattern. With the maximum directivity synthesis, no pattern control is
possible. The pattern that results must be accepted. Similarly, the unit-circle
method of Schelkunoff (see Section 9.1) provides an efficient pattern, even though
no direct pattern control occurs. Efficient here means a narrow beamwidth for a
given sidelobe level. For an efficient narrow beam pattern the array zeros are placed
on the unit circle. For d = X/2, the unit circle is traversed once (i// = 2rr) for
0 < 6 < 180. A spacing of X/4 allows the unit circle to be traversed halfway, and so
forth. Schelkunoff's method is to equally space the zeros in the applicable part of
the unit circle, \p = 4nd/X. This gives superdirectivity for d < X/2. If the zeros were
distributed over 2TT, the gain would be normal. By making minor changes in zero
locations of the superdirective array, the combination of narrow beamwidth and
low sidelobes can be improved further. This is the Chebyshev array, also described
in Section 9.1.1; directivity formulas are given there.
For superdirectivity, the correspondence between the array-pattern polynomial
and the Chebyshev polynomial must be made as follows and works only for an odd
number of elements where TV = 2Af + 1. Here a Chebyshev polynomial of order M
(with M zeros) is used. To obtain the 2M pattern zeros, the mapping starts at
z = + 1 for the main beam, then the Chebyshev ± region is traversed and mapped
onto the pattern sidelobes. At z = — 1, the Chebyshev ± 1 region is traversed again
to fill out the remaining pattern sidelobes. This scheme is valid for d</2. The
formulation of Drane is used here:
MI

(9.126)
where xn = cos (nir/M), m = 0 is the centre element and Ml = INT(M/2.)9 Ml =
INT((M + l)/2.)
This is used in double-precision form due to the partial cancellation that occurs in
superdirective arrays.
Using the formula for array directivity and the Chebyshev coefficients, direc-
tivity for arrays of 3, 5, 7 and 9 elements have been calculated, for sidelobe ratios
of 10 and 20 dB. The superdirectivity can be seen in Fig. 9.38 and 9.39 for spacing
below 0.5X, as the ordinary directivity (using the Chebyshev coefficients that are
independent of spacing) goes smoothly to zero dB at zero spacing. The figures dis-
play these calculated directivities vs. element spacing. Thus a 3-element array
offers roughly 3 dB extra directivity for small spacings, 5 elements offers roughly
5 dB. Next the disadvantages of superdirectivity will be discussed.

9.1.3.5 Efficiency, tolerance and bandwidth: The three undesirable features


that accompany superdirectivity are reduced radiation resistance (which reduces
64 Linear arrays

01 02 03 0U 05 06 07 08 09 10

Fig. 9.38 Chebyshevarray directivity, SLR = 10dB

ttn

u 01 ut us u.*» ^/^ u.3 o1^) Tlo

Fig. 9.39 Chebyshev array directivity, SLR =20dB


Linear arrays 65

efficiency), sensitive mechanical and electrical tolerances, and high Q (narrow band-
width). First simple calculations are made using Dolph-Chebyshev arrays.

Dolph-Chebyshev arrays: Because superdirectivity involves a partial cancella-


tion of the element contributions at the main beam peak, with more cancellation
for more superdirectivity, the tolerance of each element coefficient (excitation)
becomes smaller (tighter) with more superdirectivity (Uzsoky and Solymar168). A
simple calculation has been made of these effects for Dolph-Chebyshev arrays by
perturbing the centre element of an odd array, finding the tolerance to reduce the
directivity 1/2 dB. This is not expected to be sensitive to sidelobe ratio and a value
of 20 dB was used. Calculations were made for N = 3 and 5 as a function of spacing
with the results shown in Fig. 9.40. It was noted that the % error vs. d/\ curve is

Fig. 9.40 Centre-element tolerance for 0.5 dB reduction in directivity

linear up to spacings of roughly 0.1X. For N = 3 the slope is 2:1, and for TV = 5 the
slope is 4:1. Thus for TV = 3, if the element spacing is halved, the tolerances must
be 4 times tighter. The percentage tolerance for the centre element, to maintain
close to (0.5 dB) the expected directivity, is approximately:
100
% tolerance ^ —==

Q for an array (of isotropic elements) is given by (Lo et dl.105):


66 Linear arrays
N
IA
Q ~ N N
l (9.127)
I I AnAmsim[(n-m)kd]
Q, which is the inverse of fractional bandwidth, is plotted for the Dolph-Chebyshev
arrays of the previous Section in Fig. 9.41. The N = 3 array has a log-log slope of
Q vs. spacing of 4:1, while the N = 5 array has a slope of 8:1. The N = 7 and 9
arrays have even higher slopes. For small values of Q, the curve is not too accurate.

10?

105

10*-

io 2 H

10-

1
0.01 0.02 0.04 0.1 0,2
d/A
Fig. 9.41 Q of broadside arrays

Bandwidth appears to be more restrictive than tolerances; theiV = 3 array, to be


practical, requires a spacing of the order of 0.1 X or larger. And the ^ > 3 arrays are
even less forgiving. One last example should suffice to encompass the performance
of superdirective arrays. This is an aperture of X/2 wide filled with 1,3,5,7 and 9
elements of a 20 dB Dolph-Chebyshev array. In Table 9.10 are given the pertinent
values.
Linear arrays 67
Table 9.10 Dolph—Chebyshev array half-wavelength long
N Spacing Directivity Q
dB
1 _ 0 l
3 0.25 3.2 1.6
5 0.125 5.0 9 x 103
7 0.0833 6.4 6x 108
9 0.0625 7.5 2 x 1014

The currents for the last case,TV = 9, are:


52,202,865.7
-42,019,827.9
21,402,217.1
-6,306,630.6
823,815.5
In the broadside direction, the net radiating current is only 19.9, a reduction of
3 x 106 over the centre current. These results are similar to those of Yaru183 and
Jordan and Balmain,87 although his hand-calculated numbers are slightly off.
Although the Chebyshev array has an efficient pattern, other designs may offer
somewhat better power efficiency or bandwidth or tolerances. Constrained synthe-
sis techniques for this are discussed next.
Maximum directivity subject to constraints: Since a modest amount of super-
directivity may be tolerable, the directivity should be maximised with a constraint
on whichever aspect is most important: efficiency, tolerances, or bandwidth. For
example, a minimum value of efficiency may be set for an array of fixed size and
number of elements. An equivalent problem at UHF and higher frequencies, where
external noise temperature is smaller than typical preamplifier noise temperatures,
is to maximise S/N. That is, as directivity increases, efficiency increases but affects
S and N alike as long as the antenna-preamplifier system is external noise-limited.
However, when the external temperature times efficiency becomes less than the
preamplifier temperature, the S/N peak is passed (Lo et al105). Tolerances may be
constrained by fixing the total array error variance (see Section 9.1.2) and then
maximising directivity subject to these errors (Newman et al.113). Bandwidth or Q
may also be constrained; in fact, this was the first constrained superdirective optimis-
ation. Taylor158 defined a supergain ratio 3T9 which is the ratio of total (radiated plus
stored) energy in the field of the antenna to radiated energy. It is related to Q by:
JF= 1+G (9.128)
125 126
Rhodes ' maximised directivity of a line source subject to a constraint on 2f
using the eigenfunctions of the source: prolate spheroidal functions. These are par-
ticularly suited to allow a simple derivation as they are doubly orthogonal: over
infinite limits, which fits the total energy integral over all angles, and over finite
limits, which fits the radiated energy integral over real angles. Since that early work,
68 Linear arrays
more powerful constrained optimisation methods have been developed (see Section
9.1.2.) These methods can be applied to efficiency, tolerances or bandwidth, or to
any combination of these.

9.1.4 Thinned arrays

9.1.4.1 Background: Large arrays with half-wave spacing contain many ele-
ments: the associated cost, weight, etc. may be unacceptable. Also, mutual coupling
may seriously degrade scanning performance. Since half-power beamwidth is
approximately proportional to X/L, array length L cannot be reduced without
increasing beamwidth. In some cases, array size is dictated by beamwidth, but the
directivity of a filled array is not needed. Thus a thinned array can offer essentially
the same beamwidth with less directivity and fewer elements. Directivity is approxi-
mately equal to the number of elements N. Mutual coupling effects are also signifi-
cantly reduced (Agrawal and Lo2). The average sidelobe level (power) is l/N. Regu-
lar thinning produces grating lobes, but these can be suppressed by randomising*
the element spacings. There is then the question of how to select the 'random'
spacings. Most of the analytical work started with a desired pattern expressed as a
Fourier series, then devised various ways of fitting the array sum to the Fourier
series. Sandier134 expands in terms of a set of uniformly spaced arrays; Harring-
ton75 considers an array with small departures from uniform; Unz166 expands each
array term in a Bessel function series; Ishimaru83 expands in a delta-function series.
All of these approaches suffer from a common difficulty: in order to expand the
desired pattern into a Fourier or any other series, the pattern must be known in
detail. If only amplitude is known the problem is much more difficult; ideally pat-
tern amplitude and phase are given. Rarely, however, is phase specified, and for
narrow-beam, low-sidelobe arrays, the amplitude cannot even be specified. Rather
the sidelobe envelope is usually specified, or the envelope below which all sidelobes
must lie is specified. No Fourier series can be fitted to a sidelobe envelope as the
series represents the actual lobes. Thus, analytical efforts have been found to be of
little use for narrow-beam patterns. Shaped beam patterns have been synthesised by
Stutzman152 using a Legendre polynomial expansion of the array pattern, analogous
to a Gaussian quadrature.
Computer selection of element spacings has also been pursued. However, the
problem is enormously difficult as a simple calculation will show. Take an array
100 wavelengths long, and allow possible element positions every eighth wave-
length. Now take 40 elements, a 20% filled situation. There are:
/800\ IQ70

\40J "
possible patterns to be evaluated if a direct search is used. Dynamic programming
was tried as a search speedup, but since the optimum position of an element
If the array spacings are selected by a deterministic algorithm, the array can be called
aperiodic instead of random.
Linear arrays 69
depends upon the positions of elements on both sides of the element, the problem
does not meet the necessary criteria, and some empirical adjustments are needed to
get useful results. Other search speedups have proved unsatisfactory. Optimisation
algorithms using conjugate gradient methods, such as Fletcher-Powell or Fletcher-
Reeves*, will yield an optimum thinned array, say minimum sidelobe envelope for a
given array length, but are exceedingly slow. Most of the random-array computer
studies have been on small arrays (King et al.;92 Andreasen;8 Redlich;121 Stein-
berg148). Use of such arrays is limited by the type of sidelobe structure that usually
appears. High sidelobes occur at random angles, with a wide range of sidelobe
heights in between these high lobes. Thus, the computer design usually attempts to
keep all sidelobes below a fixed value. Algorithmic comparisons have been made by
Steinberg.148

9.1.4.2 Probabilistic studies: A major advance in understanding the behaviour


of random thinned arrays was made by Lo.104 He assumed isotropic elements of
unit excitation were placed randomly, giving a probability density function. From
this the probability of a narrow beam pattern with all sidelobes below a given level
was calculated by approximating the autocorrelation function by a rectangle of
width lifkjL. However, the accuracy is poor for probability (that all sidelobes are
below a specified value) values that are small, or for small arrays. A different
method was used by Agrawal and Lo2 wherein the probability of the pattern func-
tion crossing the specified sidelobe level was examined. The resulting integral equa-
tion was solved numerically by dividing the interval into many short segments, with
the probability of a crossing assumed constant over a segment. The resulting
product in the limit gives a closed-form probability result that compares very well
with Monte Carlo simulations of arrays as short as 5 wavelengths for low as well as
high probabilities. The Agrawal and Lo result is in terms of the array length L/X,
and a combined parameter which is number of elements N times power sidelobe
level:

a = — (9.129)
SLR
Here the sidelobe ratio SLR is in power. The result is:
p = [l-e-*]e-(2L/\)*j7rot/3exp(-a) (9.130)

Thus sidelobe level and number of elements can be directly traded. Fig. 9.42 shows
P vs. a for L/X = 10, 30, 100, 300, 1000. For example, mSLR of 20dB is achieved
with an 80% probability for L = 1000X withTV = 1028; or an average element spac-
ing of one wavelength. For large N the formula may be simplified (Lo104):
p = [ 1 - <,-<*]6L/\ (9.131)

This gives a reasonably close fit for P>0.5 and L/X> 100. Fig. 9.43 shows the

These are in most computer scientific subroutine libraries.


70 Linear arrays

approximate relationship for P = 0.8, 0.9 and 0.99 vs. L/\. The quantity plotted is
\0\ogN —SLR and again these results are only valid for large arrays. All of the
results are for an unsymmetrical array. If the array is symmetric about its centre
some degrees of freedom are lost, and the sidelobes are higher.

09-

08-

07

0 6

!5 0 5

0-4.

03

01

U
N/SLR

Fig. 9.42 Probability of random-array sidelobe ratio greater than SLR

In using the curves of Fig. 9.43, care must be taken that the resulting TV is not
greater than that of a filled array, 2L/\. For example, a filled array with/, = 100X
has 7V= 200 (half-wave spacing) which gives a = 9.44 for SLR = 13.26 dB. From
the curve the probability is near unity, which means that a 100X aperture can
achieve an SLR fo 13.26 dB or better. However, for L = 30X, N = 60 the value of
a for that SLR is 2.83, which from the curve gives a probability near zero. This
means that a thinned 30A aperture will probably have sidelobes worse than — 13 dB.
Linear arrays 71
Thus, for each aperture length there is a sidelobe ratio that is difficult to exceed,
with longer apertures yielding better SLR capability. For example, for 80% prob-
ability, Table 9.11 gives these values. Thus, for reasonably low sidelobes very large
arrays are required to achieve a significant thinning. But these very large arrays have
very many sidelobes; only a small fraction of these sidelobes need to be at the
envelope level to greatly reduce directivity and beam efficiency. This is expected
since directivity is roughly proportional to the number of elements. Beam
efficiency is important for radiometer and power-transfer situations. A tapered
sidelobe envelope will reduce the sidelobe power and increase beam efficiency. This
is especially important for planar arrays since there are so many more sidelobes at
large angles from broadside.

Table 9.11 Upper limit of sidelobe ratio


L/\ SLR
dB
10 5.7
30 9.6
100 14.1
300 18.2
1000 22.9
3000 27.2

P = 0-99

dB

09

08

100 1000

L/X

Fig. 9.43 10logN - SLR for a thinned array


72 Linear arrays
9.1 A3 Non-hotropic elements: The analyses and results above are for thinned
arrays of isotropic elements. Half-wave elements will tend to reduce far-out side-
lobes in one plane due to the element pattern. Often it is advantageous to use ele-
ments larger than half-wave, and these further reduce far-out sidelobes. Steinberg148
has shown that the effect of element directivity is to reduce the effective array
length used in eqn. 9.130. His result is

(g 1321

This is only valid for W/X > \, due to approximations used in evaluating an integral.

Table 9.12 Thinned array with directive


elements: L = 1000X,
P=0.8,SLR= 15dB
W/X P Filling factor
Isotropic 0.8 0.163
0.5 0.87 0.163
2 O97 0.651

An example will show the effect of element directivity. TakeL/X = 1000, P = 80%
and SLR — 15 dB. Then a thinned array of isotropic elements needs 325 elements.
Keeping the SLR and the number of elements fixed, consider half-wave dipole-
type elements (W/X = ^) and 2X sub-arrays. Table 9.12 gives the probabilities of
sidelobe ratio of 15 dB or greater. As expected then, directive elements reduce the
sidelobes.

9.1.4.4 Space tapering: A special kind of thinned array uses variable element
spacing to produce an equivalent amplitude taper. The goal is to: (a) produce a side-
lobe envelope that tapers down, and (b) produce lower initial sidelobes (Willey177).
As in all thinned arrays the number of array elements must be large. It is particu-
larly attractive in a distributed amplifier array where each element has its own
transmitter/receiver module. Since it is much easier to make all modules alike, space
tapering allows lower sidelobes to be realised, with all elements equally excited.
Experience has shown, however, that the sidelobes do not fall off quite as fast as
predicted by theory. Also an occasional high sidelobe is observed. It may be
necessary to make adjustments to the spacing or to the filling ratio.
Some recent calculations show the degree of thinning vs. sidelobe level. The
Taylor one-parameter distribution (see Section 9.1.1.3) is used. The number of
elements relative to the X/2 spacing filled array number No is:
Linear arrays 73
This can be integrated (using eqns. 11.4.10 and 10.2.13 of Reference 1) to give:
N= (9J34)
~^dm
Table 9.13 gives the values and also shows the relative beamwidths (taken from the
Taylor space factor). Even for SLR = 40dB, the thinning is less than half. So,
although space tapering does allow simulation of tapered amplitude distributions,
the percentage of elements saved is modest. For rectangular arrays, the values for
each co-ordinate are combined.

Table 9.13 Space tapered linear array


SLR N/No Wo
dB
13.26 1 1
15 0.9133 1.042
20 0.7562 1.156
25 0.6647 1.260
30 0.6030 1.355
35 0.5577 1.443
40 0.5223 1.524

9.1.5 Array tolerance and limits

9.1.3d Tolerance in uniformly spaced arrays: Errors in array element excita-


tion, location and orientation will affect the array pattern, directivity and other
performance parameters. Effects of errors in limiting directivity, and optimisation
of directivity subject to a tolerance constraint, are covered in Sections 9.1.2 and
9.1.3. Of interest here are the direct pattern and directivity effects. Bias errors are
both relatively easy to understand and difficult to categorise; thus only random
errors will be analysed. All random errors will have an assumed normal (Gaussian)
distribution with zero mean and variance a2. Linear arrays are particularly appro-
priate for analysis as they are simple, yet the principles and results apply with only
minor changes to planar arrays. Isotropic elements will be considered. The space
factor is written as:

fi>(«) = I Aneimm (9.135)


n

Now let the amplitude error be An and the phase error be 6 n ; the coefficient with
error is:
An(l + K)eiSn
The power pattern is now:
74 Linear arrays

E\B) = 1 1 ^ n ^ ( l + A n ) ( l + A
(9.136)
Under the assumption of zero means, and by using the Central Limit Theorem, that
the mean of the sum of independent variables is the sum of the means, the average
pattern with errors becomes:

^ ( 0 ) = £3(0) + £ - (9.137)
where E% is the error free pattern, Go is the directivity, and
o 2 = o2A + o28 (9.138)
This interesting result, first developed by Ruze,131'132 shows that on the average a
small constant amount is added to the sidelobes, and that this effect is less for long
arrays (high directivity). Also for low-sidelobe designs the effect is more pro-
nounced as addition of a small error field affects a low sidelobe more than it would
a higher sidelobe. The main beam peak is reduced a small amount. The probability
density function PDF of the field is a modified Rayleigh (Bennett14), because the
field at any angle is the vector sum of the no-error field and the x and y error
components.

(9,39)

The variance o2E is o2/G.


Although the mean sidelobe performance has been simply described, the prob-
ability of individual sidelobes exceeding the mean is of importance. This is found
by integrating the PDF from the specified sidelobe level (SLL) to infinity
(Rondinelli129), for the ith sidelobe:

P(Et>SLL) = J sLL PDFdE (9.140)


By change of variable this can be written in terms of the Marcum Q Function
(Marcum108):
P(Et>SLL) = Q(AhB) (9.141)
where Ai — \[lMiJoE and B = y/2SLL/oE. The Q function is defined as (Mar-
cum108):
Q(A9B) = j ~ t exp f ^ y ^ j Io(At)dt (9.142)
The expression for probability, however, is only for one specified sidelobe to
exceed SLL. The probability that any sidelobe exceeds SLL is:

P = I " ft [l-Q(An,B)] (9.143)


Linear arrays 75

In this expression An represents the peak of the unperturbed nth sidelobe, and the
product is over the desired range of sidelobes. Using a recursive algorithm developed
by McGee,110 values of probability have been calculated for arrays of 20 and 100
elements; these numbers represent the range of many actual arrays that are being
used. Figs. 9.44 and 9.45 are for uniform excitation and show probability of side-
lobes from 10° to 90° exceeding — 15 dB, and of sidelobes from 30° to 90° exceed-
ing — 20 dB. 10° is between the first and second sidelobes; the latter has a nominal
level of —17.7 dB. And 30° is between the fourth and fifth sidelobes, with the
latter nominally at —23.6 dB.
- 2 0 , dB beyond fourth sidelobe
1.0 -,
—15 dB, beyond first sidelobe

0.8-

0.6 -

0.4 -

0.2 -

Fig. 9.44 Probability of sidelobes above stated level


Uniform, N = 20

As expected the larger array is less sensitive to errors. Arrays of 20 and 100 ele-
ments using a Taylor one-parameter 25 dB distribution have been similarly calcu-
lated, with the results shown in Figs. 9.46 and 9.47. Figs. 9.48 and 9.49 are for a
low-sidelobe 40 dB Taylor one-parameter space factor. Again, the cases are for side-
lobes past the first and past the fourth. Note scale changes as the design SLR
changes. In each case the dB level chosen is roughly the same number of decibels
above the nominal second sidelobe level, and similarly for the level above the fifth
sidelobe level, thus allowing the effect of initial sidelobe ratio to be observed.
This theory assumes that the perturbed sidelobe peak position is unchanged,
which is valid for small to modest errors. Thus, to calculate probability of sidelobes
exceeding a given level, only the levels of the unperturbed sidelobes need to be
known. The examples given here are not sufficiently extensive to cover all cases
that might arise. However, the Q function is easily put into subroutine form and the
calculation of probability is quite simple.
Directivity reduction is found by inserting the modified power pattern into the
directivity expression. The result, for small errors, is:
76 Linear arrays

Go
/~% (9.144)
2
l+o
Thus a one sigma phase error of 30° gives a 1.05 dB directivity loss. A one sigma
amplitude error of 1 dB yields a 1 dB directivity loss. Note that directivity loss, to
first order, is independent of array size, directivity or sidelobe level.

10-

1-4

F ig. 9.45 Probability of sidelobes above sta ted level


Uniform, N= 100

Other errors than those in amplitude and phase can be simply related to ampli-
tude and phase if they appear singly. Combinations of various types of errors can
be analysed if the relative values of the variances of each error are known
(Elliott44).

9.1.5.2 Directivity limits due to sidelobes: The relation of achievable directivity


to average sidelobe level implies that space factors with sidelobe envelopes above
1/u may incur a directivity limit. For example, a Taylor 'ideal' or Dolph-Chebyshev
space factor will contain an amount of energy in the equal-level sidelobes that varies
only slightly with L/X for long sources. Yet the main beam energy decreases
inversely with L/X. When the main beam and sidelobe energies become comparable
the directivity is approaching a limit. This has been calculated for the Taylor 'ideal'
and n sources (Hansen67). Key to the directivity calculations is an integration of the
Linear arrays 77

0.30

Probability of sidelobes above stated level


25 dB Taylor, N = 2 0

I.O-i

0.25 0.30

Probability of sidelobes above stated level


25 dB Taylor, N = 100
78 Linear arrays

- 4 8 dB
beyond fourth
l.O-i

' -42 dB
beyond first
0.8-

0.6-

/ /

0.4- /

u
0.2-

/ /

0-
0.01 0.02 0.03 0.04 0.05
O"

Fig. 9.48 Probability of sidelobes above stated level


40 dB Taylor, N = 2 0

I.Oi

0.8-

0.6-

0.A-

0.2-

0 0.01 0.02 0.03 0.04 0.05

Fig. 9.49 Probability of sidelobes above stated level


40 dB Taylor, N= 100
Linear arrays 79
pattern over the main beam and all sidelobes. Alternatively, the aperture distribu-
tion could be integrated. In the references cited numerical integration was used for
the Taylor 'ideal' space factor, while the Taylor n space factor was integrated by
approximating each sidelobe by a sinusoid of the proper amplitude. An alternate
scheme could use the directivity formulas given in Section 9.1.1.3 for Dolph-
Chebyshev and Taylorftdistributions. Fig. 9.50 shows the directivity limitations on
the 'ideal' space factor. For the n space factors, Figs. 9.51 and 9.52 show the limita-
tions for half and for a quarter of the sidelobes of equal level. Longer line sources
obviously must employ smaller n to avoid directivity loss. No loss is experienced by
the Taylor one-parameter space factor, since the sidelobe envelope already decays
as l/u.

1.L

uniform
1.0 -

taytor ideal line


source sidelobe ratio
0.8-
^ 35 dB

\30dB
0.6 -

^ \
0.4 - \25dB
\

^ \
0.2 - 20 dB

1 1 —I ! I
0- 400 600 1000
40 60 80 100 200

Fig. 9.50 Gain limit of ideal source

9.2 Array feeds

Linear array feeds in general are either series or shunt, which refers to the geo-
metry rather than to an equivalent circuit. Series arrays, sometimes called line
arrays, have the elements disposed along a transmission line. The line can be a
standing-wave (resonant) feed, described in Section 9.2.1, or a travelling-wave feed,
described in Section 9.2.2. Some other aspects of line feeds are discussed in Section
9.2.3, along with shunt feeds where each element is fed independently, either by a
divider network (corporate feed) or from its own receiver/transmitter module
(distributed array). Most of the discussion is centred on dipole or slot elements.
80 Linear arrays

12-1

10-

sidelobe ratio
35dB
0-8-

06-

0 4-

02-

20 40 60 80 100 200 400 600 1000


L/X

Fig. 9.51 Gain limit of n source


N = M/2, Half of sidelobes of equal level

12-,

10-

sidelobe ratio
35dB
08-

0 6-

04-

02-

10 20 40 60 80 100 200 400 600 1000


L/X

Fig. 9.52 Gain limit of n source


N = M/4, quarter of sidelobes of equal level
Linear arrays 81

Microstrip patch elements are used both for resonant and for TW arrays; see Section
9.3.4. Another type of TW array is the serrated guide with closely spaced transverse
broadwall slots (Hyneman81). This is a type of leaky-wave antenna (Tamir155).
To scan an array each element must be excited with the proper progressive
phase; a phase shifter or phaser is used. However, these devices are not covered here
(see Ince and Temme82).
Quantisation effects on pattern and gain are discussed in Section 9.2.4, where
the array is quantised into sub-arrays, or where the phasers are quantised.

9.2.1 Resonant arrays

9.2.1.1 Impedance of resonant arrays: A resonant array or resonant feed has


elements spaced at half-guide wavelength intervals; typically the end is shorted so
that a standing wave is set up along the guide. Input power can be applied to the
other end, or along the guide, in which case both ends are shorted. Each element
provides a small reflection, and at the centre frequency these reflections add up. A
precise theory should include all multiple reflections at any frequency of interest
and should include the change of element admittances with frequency. This can be
done using transmission line theory. If the array excitation is uniform, then a trans-
mission line with periodic loads represents the resonant feed. A closed-form solu-
tion can be obtained even if waveguide loss is included. Here the periodic loads are
the radiating or coupling elements, all with the same admittance. The periodically
loaded line is equivalent to a uniform line with complex propagation constant
(because of the slot conductances) and complex characteristic impedance
(Slater;140 Altmann5). However, the equations are too complicated to permit extrac-
tion of simple results. More important, tapered amplitude excitations cannot be
handled this way.
For general excitations a numerical approach is satisfactory. The resonant feed is
represented by a cascade of transmission-line sections with an admittance in shunt
at each junction. The overall input admittance is then that looking into this cas-
cade, with a short at the far end. A-B-C-D matrices are ideal for this formulation
as the matrix of a cascade is simply the product of the matrices. The prototype sec-
tion is a transmission line of length / with a load admittance Yn, see Fig. 9.53.

Y0 = ' % Yn
o o 6
pi
Fig. 9.53 Prototype sect/on

Calling C = cos j3/, S = sin j3/, the A-B-C-D matrix of the section (Altmann5) is:
C jS\(l 0\ lC + jYnS jS
(9.145)
cj\rn 1/ \Ync+js c;
82 Linear arrays
The overall matrix is then the product of the prototype matrices;one is used for each
element in the array, starting at the feed end. A shorted quarter-wave section with
C2 = cosj3//2 and large Ys is added at the end, giving:

N jS2
- (9.146)
c D Ysc2+jS2 c2

Here C is used both as part of the A-B-C-D matrix and also as a trig term, but the
usage should be clear. Either reflection coefficient or input impedance can be calcu-
lated from A-B-C-D. The latter is:
A +B
(9.147)

3-

to

2-

102 1-0A 106 108 110


f/fn

Fig. 9.54 VSWR of uniform resonant array


N= 20

and from this the VSWR is obtained. Fig. 9.54 is for an end-fed uniform array of 20
elements, showing VSWR vs. normalised frequency. Only f/fo> 1 is shown as the
curve is symmetric. The nearly linear part of this curve, with a similar curve for a
25 dB Taylor one-parameter distribution are in Fig. 9.55 plotted against NB, where
N is the number of elements and B is the total percentage bandwidth. The Taylor
array is half of a symmetric array; centre feed is assumed with each half matched.
Linear arrays 83
The curve is for N elements in the half array. From these the values of NB for
VSWR = 2 are 66 for the uniform case and 50 for the 25 dB Taylor. Tapered distri-
butions have narrower bandwidth when the strongly coupled elements are near a
feed point. This can be seen by visualising the array on a Smith chart. Starting near
the top (slightly off resonance) the first load moves the point along a constant B
line toward the centre. Next the line length gives a rotation that is nearly 360°.
Then the next load moves to a larger B, etc.

3.0

25 dB Taylor
2.5 -

2.0 -

a:

3; 1.5 -

1.0
20 60 80

Fig. 9.55 VSWR of half of resonant array; number of elements times % bandwidth

When a uniform excitation is used, the largest constants movements (for small G)
are nearly radial, which moves the admittance toward match. The last movements,
which are not radial, are small on the chart. In contrast with a tapered distribution
the small G occur in the radial movement region, while the large G values (near the
feed) move to a larger B. Thus the real part of impedance is closer to unity, but the
reactance is larger than that for uniform excitation. For all distributions there is a
negligible change in NB withiV.
Returning to Fig. 9.54, it is interesting to note that the VSWR oscillates with
frequency for wider bandwidths. A simple physical explanation of this can be
developed using simplified theory developed by Watson.173 He set up a finite
product of factors, where each factor contained a length of transmission line and a
shunt load admittance. By neglecting multiple reflections between loads, the
product could be approximated by a sum. His often quoted result is:
N
. N

T = ^r N
r- 1 1 ^N (9.148)
2+ Yne J2n(3d
I I
n=l
84 Linear arrays
At centre frequency 0d = n and the formula reduces to the simple form:
N

(9-149)

From this arises the simple rule that at centre frequency the sum of the slot sus-
ceptances should be zero and the sum of slot admittances should be unity. Fry and
GowardS9 have plotted eqn. 9.148; the VSWR rises with frequency to a normalised
frequency just above 1 + 1/2N (for uniform illumination) and for higher frequen-
cies the-VSWR oscillates about 2. From this simple theory the near linear region
gives NB = 89 for VSWR = 2 and N = 20. Note that this result is more generous
than the exact result quoted earlier. When the excitation is uniform, Yn = l/N and
the series can be summed exactly, giving:

1 KN-i

r . ^ ( 9 . 1 5 o,

Nsmpl
The sin factor is recognised as the space factor of aniV-element array, and this fac-
tor produces the oscillations in F and in VSWR as j3/ changes. The oscillation is
about F = 1/3 or VSWR = 2 since the centre of the sin factor oscillations is zero.
An even more simple derivation observes that at a frequency where the last ele-
ment is n out of phase, N$l = Nn + n/2 or/// 0 = 1 ± 1/2N, which gives a product of
NB = 100 (full bandwidth). From eqn. 9.150 this corresponds to T = 0.3475 and
VSWR =2.07.
All of the results given so far are too optimistic. When the changing element
impedance (both self and mutual) with frequency is included, the bandwidth drops
significantly, perhaps to 2/3 or less. With waveguide broad-wall shunt slot or dipole
elements, the element impedance can be included in the calculation. For waveguide
edge slots it is necessary to measure the actual bandwidth. Another practical factor
is the beam splitting that occurs at roughly twice the NB for VSWR = 2. The two
travelling waves that at centre frequency combine to give a standing wave are now
of unequal amplitude and out of phase, so that one travelling wave produces a beam
left of centre and the other produces a beam right of centre (Kummer98). However,
this effect is not noticed for bandwidths that keep VSWR < 2.

9.2.1.2 Resonant-slot Array Design: A resonant array, sometimes called a


'standing-wave array', consists of radiating elements located along a feed transmis-
sion line at half wavelength intervals. The excitations are alternately in phase and
out of phase; the latter are usually corrected by reversing the terminals of every
other dipole connected to a two-wire feed line, or by reversing the sense of coupling
of slots in a waveguide feed. Sense reversal for slots is accomplished by moving
Linear arrays 85

broad-wall displaced slots across the wall centre-line, while inclined broad wall or
edge slots are reversed in inclination. As a result, all elements are excited in phase
and principal radiation occurs in a broadside direction. Resonant arrays are most
often constructed of waveguide slots due to the precision with which such slots can
be milled. Due to the half-wave element spacing (in feed guide wavelengths) the
input admittance is simply the sum of the slot admittances plus the end admittan-
ces. A short is properly located at the end so that the input admittance depends
only on the slot admittances. Such a linear resonant array may be fed from one
end, but more often both ends are shorted, with a cross-guide coupler used as feed.
The flat plate array of Chapter 10 is of this type. The resonant nature of the feed
introduces bandwidth constraints.
The slot amplitude distribution can be obtained from the techniques of Section
9.1; next is the question of determining the set of slot offsets (or angles) and
lengths, assuming that slot width and waveguide dimensions have already been
selected. However, an important factor makes the simple calculation of slot para-
meters using the universal slot data of Section 9.3 incorrect. This is the ubiquitous
mutual coupling which has a significant external component, due to coupling
between slots in the region outside the waveguide, and may have a significant
internal component due to slot coupling inside the guide. The latter is implemented
via the excitation of higher modes which, though evanescent, have small decay from
a slot to its neighbours. Internal coupling is severe for edge slots and is second order
for broad wall slots. Due then to mutual coupling, which in practice couples each
slot to several slots on each side, the slot-array design problem must be solved
simultaneously instead of slot by slot.
Two relationships allow such a solution. First the sum of slot admittances,
including mutual admittances, equals the specified input admittance. For a single-
line array this is usually 1 + / 0 , but for planar arrays the real part may be different
from 1. Secondly, all mode voltages in the waveguide at each slot are equal. How
then to incorporate mutual coupling? Broad-wall slots can be replaced by Babinet
equivalent strip dipoles if the waveguide face is assumed to be a flat ground plane.
The remainder of the discussion applies to broad-wall slots. Most slots are narrow so
that the strip dipoles can be replaced by cylindrical dipoles of diameter half the
strip (and slot) width (Tai153). Then Carter's 1932 mutual impedance formulas
using zero-order current distributions on the dipoles can be used. A convenient
formulation of these for computers has been made by Hansen and Brunner.73 Use
of more accurate current distributions is not warranted as the errors introduced via
the ground-plane approximation are more significant. In designing a resonant array
of dipoles, the Carter mutual impedances, suitably weighted by the excitations, are
simply added to the dipole self impedances. Slot self admittance, however, includes
the effects of energy storage inside the guide in the vicinity of the slot, and thus the
waveguide slot is not the Babinet equivalent of the strip dipole for self-admittance
purposes. A clever solution to this dilemma was produced by Elliott and Kurtz.49
Slot self-admittance is determined by measurement or by calculation from the uni-
veral slot data. Then it is inverted to slot impedance and the mutual-impedance
86 Linear arrays
terms from the strip dipoles are added. This is again inverted to produce active slot
admittance. Turning now to the design equations for longitudinal displaced shunt
slots, slot admittance, using the modified Stevenson result of Section 9.3, and the
mutual impedance process just described, becomes:

(9.151)

Here Zn is the combined inverse of waveguide slot admittance and dipole mutual
impedances; Yn is the active admittance of the nth slot and the coefficient Cn is:

Cn = s i n — (cosjM n -cos« n ) (9.152)


a
From network equations applied to an array of dipoles, the combined impedance
can be written as:
iSOa/bCj f F A ,
, V

In this formulation the series does not include n = m as the waveguide slot self
admittance Y® has been written separately. Vn is the slot voltage and Znm is the
mutual admittance between the nth and rath dipoles. Thus, if the slot offsets (or
angles) and lengths are known, eqn. 9.151 with the help of eqns. 9.152 and 9.153
can be used to obtain the active slot admittances. Then the sum should be:
N y
£ -n = i + / o (9.154)
n-l ^0
This equation is complex, and hence gives two real equations. Additional equations
arise from the equality of mode voltages mentioned earlier. Mode voltage vn at the
nth slot is:

When each mode voltage is equated to the next, vt = z>2, v2 = v3,... pn_ t = vn, a set
of 2N — 2 real equations is produced. Thus, there are now 2N total equations relat-

Fig. 9.56 Resonant slot array (courtesy: Antenna Dept., Radar Systems Group, Hughes,
El Segundo)
Linear arrays 87

ing the 2N unknowns (slot offsets and lengths), so the problem is, as expected,
determinate. Unfortunately the equations are nonlinear and transcendental, so an
iterative solution must be employed. A satisfactory solution uses the Newton-
Raphson method (Stark145). Here a gradient matrix is formed, of derivatives of the
RHS of each equation, for each variable. This matrix is calculated using the vector
of variables; these are a set of 2N real simultaneous equations whose solution is the
delta vector to be added to the vector of unknowns. Through iteration a set of un-
knowns is found that satisfies the mode-voltage and input-admittance equations. A
starting vector must be provided initially. Convergence is best indicated by the
gradients becoming small. This technique has been used with success for the design
of a variety of linear and planar slot arrays. Fig. 9.56 shows a typical waveguide
resonant array.

9.2.2 Travelling Wave A nays


A TW array consists of a linear array of radiating elements disposed along a trans-
mission line. The transmission line is excited at one end; each element along the line
couples a portion of the energy; after the last element the small remaining power is
dissipated into a load. To avoid reflected-wave reinforcement, element spacing is
not half wave. Thus, the TW array is non-resonant. The spacing makes the array
main beam move with frequency. Elements near the feed must be coupled lightly,
while elements near the load must be heavily coupled. Given a desired array excita-
tion it is necessary to solve a transmission-line problem to determine the coupling
distribution.

9.2.2.1 Single-beam condition and frequency squint: The feed line may be
TEM or dispersive, and may have phase velocity above (waveguide), equal to, or
below (dielectric loaded coax) that of free space. For what range of spacings will
only one beam exist? Two examples are representative. First is a TEM line, where
the element couplings are n out of phase. Why out-of-phase coupling will be
explained later. With 6 measured from broadside and d element spacing as usual,
the interelement phase shift needed for a main beam at 60 is kd sin d0. Phase along
the feed line is kd + rr, where the IT is the out-of-phase term. Thus the phase
equation is:
kdsindo + 2nn = kd-n (9.156)
Each n > 0 represents a real beam, with the nth beam starting at d/\ = (n 4-§)/2.
Fig. 9.57 shows beam angles vs. d/X for the first three beams. No beam with com-
plete addition exists for d/X<0.25, and for d/X= 0.75 two beams exist. So the
useful range is 0.25 <d/X< 0.75. Either forward squint (d/\> 0.5) or backward
squint (d/\ < 0.5) is possible. Beam-squint change with frequency is indicated by
the slope of the n = 0 line. Spacing of 0.5X represents a special case of broadside
radiation. However, this spacing allows all element admittances to add together,
with the resulting narrow-band performance of the standing-wave array of Section
88 Linear arrays
9.2.1. Now it will be evident why TW arrays with in-phase coupling are undesirable.
The phase equation in this case is:
= kd (9.157)

0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.

Fig. 9.57 TW array beam angle vs. spacing

This allows a forward fire (endflre) main beam for n = 0. (sin0o = 1) a* all spacings.
Although beams for n = 1,2 etc. can be positioned at a desired angle, the endflre
beam is always present.
The second example is that of a waveguide array with out-of-phase coupling.
Again, the in-phase coupling allows a beam for all spacings, at sin#o = &/k- F ° r the
out-of-phase case the phase equation is:
0d-ir (9.158)
Now both plus and minus values of n give real beams. Fig. 9.58 shows beam
position vs. d/X for the first three beams. The graph is plotted for P/k = 0.6. For
P/k = 0.5 both n = ± 1 modes start at d/X = 1. Since waveguides are usually
operated with higher ($/k, the range of operations is:

0/*>O.5 (9.159)
1+P/k X 1 + PII
For the example chosen, the n = — \ beam starts at d/X = 1.25 where the n = 1
beam starts at d/X = 0.9375. Not shown is the n — 2 beam starting at d/X = 1.563.
The Figure gives a picture of the mode appearance but is not useful for design since
it is for P/k = 0.6 only. Fig. 9.59 plots beam angle vs. spacing for various values of
P/k: P/k = 0.5(0.1)0.9. Using this graph the tradeoff of p/k and d/X to give a fixed
beam angle may be made.
Linear arrays 89
1-0-,
n=-1

Fig. 9.58 TW array beam angle vs. spacing


p/k = 0.6

08

Fig. 9.59 TW array beam angle vs. spacing; single-beam region


90 Linear arrays

Change of beam angle with frequency is important, as can be seen from the slope
of the curves. An explicit formula for slope is useful. Taking the derivative of
sin QQ w.r.t. frequency, and multiplying by frequency to normalise the slope, gives:
tfsinflo X2Ac , 1
(9.160)
df p/k 2d/X
This contains three interrelated variables, but fortunately it simplifies to:
d sin 60 1
— sin #o (9.161)
df
Normalised slope is plotted vs. sin#0 in Fig. 9.60. As expected, P/k nearer to unity
gives lower slopes, as do larger values of sin d0. The curves stop at the angle where a
second beam emerges; thus the practical design range is contained in these curves.
The beam angle, waveguide P/k, element spacing and normalised slope can all be
traded off for an optimum design given a particular application.

- 1 0 -0-8 -0-6 -CU -0-2 0 02 0U 06 08 10


sine 0

Fig. 9.60 Beam-angle change with frequency

9.2.2.2 Element conductance calculation: An exact calculation of the coupling


in a travelling wave array would use A-B-C-D matrices to combine each segment
of waveguide (between slots) and one slot admittance with other segments, finding
Linear arrays 91
the overall input admittance, as is done for resonant arrays. However, since TW
arrays usually have many elements, each coupling is then small, and since the
element admittances add in a pseudo-random manner due to the irrational spacing,
it is usually sufficient to assume that the array is matched at any point. That is at
each element the power divides, with part producing the element excitation and the
remainder going on to the next element after some waveguide attenuation. Let the
desired array-excitation power distribution be proportional to Fn, n = 1 to N. Call
the fraction of input power that is dissipated in the end load L:
p
load
T — rn i/n\
L
~ P (9.162)
and call the conductance distribution along the waveguide Gn. If waveguide loss is
included, the factor s representing power loss between slots is appropriate:
s = exp(-2cd) (9.163)

F F F F
1 2 N-1 N
t f ft

Fig. 9.61 Travelling-wave array coupling

Interelement spacing is d and the attenuation coefficient a can be found in standard


texts on electromagnetic theory. Using the sketch of Fig. 9.61, the last coupling is
written first, then the next, etc.

in5 l (9.165)

The fraction of input power delivered to the load isL:

L =p~ (9.166)
The input power is the load power plus the sum of coupled powers, with each
multiplied by the appropriate number of segment loss factors s~l\

Pin = £ Fmsl-m+PlSl~N (9.167)


92 Linear arrays
The last two equations can be used to eliminate Pim givingPt:

Pi = T-~^N I Fm*-m (9-168)


1 LS m=l
This is used in the expression for Gn to yield the conductance equation, after some
rearrangement:

Gn = sr^^ s (9.169)
p «-
m-\
For negligible loss, of course, s = 1. Loss generally needs including only when it is
large (e.g. at 94 Ghz) or when the loss between elements is comparable to the small-
est conductance. Usually the Fn excitation coefficients are obtained from a distri-
bution such as Taylor one-parameter or Taylor/!, and the scale factor of theFn will
make Pin ¥= 1. Note that at each slot the fraction of power coupled out is
G/(l + G). The continuous version of eqn. 9.169 is often seen (Kummer98).

G(X) = - r — ££>— (9.170)


-—j [F(y)+R(y)]dy- [F(y)+R(y)]dy
J
1 —L -L/2 "-L/2

Here R(y) is the loss per unit length of guide. However, as might be expected,
sampling the continuous result to get slot coupling is less accurate than using the
sum formula (eqn. 9.169). For small arrays, or for low sidelobe arrays, it is import-
ant to use this more exact equation.
Fig. 9.62 shows the conductance values for a 29-element uniformly excited
array, for load power percentages of 5, 10 and 25%. It may be seen that the 5 and
10% curves go off scale, while the 25% curve peaks just above 0.1. It is always
necessary to keep all conductances small; 0.1 is a good maximum. Not only are
larger values often difficult to realise, but the impedance mismatch effect may no
longer be negligible. Larger arrays, of course, will have lower conductances per slot.
The curves foriV = 29 may be converted to general-purpose curves (approximately)
by multiplying values by N/2 = 14.5. Conductance values for different N are then
obtained by dividing by N/2. However, for precise results, a new calculation should
be made after the number of slots is established based on the approximate scaling
just described.
Figs. 9.63 and 9.64 give conductances for the Taylor one-parameter distribution
and for the Taylor n distribution (n = 5). For both figures N = 29 and SLR =
25 dB. Note that the heavy taper of the one-parameter distribution gives curves
with a single peak, whereas the n distribution, which has a higher pedestal (see Sec-
tion 9.1), tends to rise again near the load, especially for small load power. Since all
the curves for 5 and 10% load power have peak above 0.1, it is in practice necessary
to either increase load power, thereby reducing gain, or to increase the number of
slots. Again the curves can be scaled approximately using N/2.
Linear arrays 93
9.2.2.3 TW slot array design: Without the effects of mutual coupling the design
of a travelling-wave slotted waveguide array is simple: a table of conductances is
obtained (see previous section) and the universal slot curves of Section 9.3 are used
to find length and position of each slot to give the proper resonant conductances.

0.28

0.2-4-

0.05/ /0.1

0.20-

S 0.16-
c
o

o
8 0.12-1
P, = 0.25

0.08-

0.04-

17 21 25 29
slot number

Fig. 9.62 Slot-conductance values; uniform excitation


N =29

With mutual coupling the slots interact strongly so that it is necessary to consider
the entire array at once. This has been successfully done for broad-wall shunt (dis-
placed) slots by Elliott,47 and the procedure can be applied equally well for any
array in which mutual coupling between elements can be accurately computed. A
salient exception is the inclined-edge slot waveguide array, as the replacement of
slots by Babinet equivalent dipoles for the computation of mutual impedance is not
sufficiently accurate.
The design procedure for travelling-wave arrays is similar to, but more compli-
cated than, that for resonant arrays. As is the case there, the mode voltage at each
slot is proportional to the slot voltage divided by the active impedance:
94 Linear arrays

(9.171)

Unlike the resonant array, the mode voltages are unequal, since the slots are spaced
an irrational number of guide wavelengths apart. Notation is that of Section 9.2.1,
except the mode voltages are subscripted here. In fact, the nth mode voltage is
related to the n — 1st mode voltage and the n — 1st admittance, where elements are
numbered starting at the load. This relation is:

= cos/H sin (9.172)


-l

028

P, =005
024

13 17 25 29
slot number

Fig. 9.63 Slot-conductance values; Taylor one-parameter


SLR = 25dB,A/=29

Yn is defined below, and d is the slot spacing along the guide. Combined with eqn.
9.171 an equation is obtained for the ratio of two successive slot voltages:

7£~ COS/3J + / - (9.173)


V , C YA /
v 1
n-1 ^n n
Linear arrays 95
Since the slot voltage is specified, if the fractional power into the load L, and all the
slot lengths and offsets are known, eqn. 9.173 represents IN — 2 equations that
must be satisfied. That is, using the correct slot parameters gives the desired slot
voltage ratios. A perturbation technique is then used with starting values of slot
lengths and offsets. However, the iV-element array has IN unknowns; the remaining
two equations come from the input admittance. Like the resonant array it should
be 1 4-/0. However, input admittance here is not just a simple sum of slot active
admittances; rather, a transmission-line concatenation must be used. Calling Yn the
guide admittance seen at the nth slot (including the guide admittance beyond the
slot and the slot active admittance), it becomes:
Yn Y£ Yn _j cosfid4- jG0 sin jkl
0 ^0

0-28-j

024-

P, =0 05

13 17 29
slot number

Fig. 9.64 Slot-conductance values; Taylor n — 5


SLR = 25dB,/V=29

Starting at the slot nearest to the load where Yx = Y* + Go, eqn. 9.174 is applied
successively until the input admittance is obtained. This should be 1 + / 0 if the cor-
rect slot lengths and offsets were chosen. These two real equations, with the 2N — 2
96 Linear arrays

mode voltage equations give the IN real equations necessary to allow a definitive
problem. These also can be solved iteratively using a Newton-Raphson solution.
Fig. 9.65 shows a typical slot array.

Fig. 9.65 Travelling-wave edge slot array {courtesy: Antenna Dept, Radar Systems Group,
Hughes, El Segundo)

9.2.2.4 Fixed squint feed: As shown at the beginning of Section 9.2.2, the
beam angle changes with frequency because of the series nature of the feed. This
beam scan can be removed by use of an equal-path-length feed, as sketched in Fig.
9.66, (Cheston and Frank28). Unfortunately, much of the simplicity and low cost of
the series TW feed is lost. The series feed slots in this scheme do not radiate
directly, but each feeds a waveguide. Radiating slots or horns are arranged at the
waveguide end. The geometry is arranged so that the waveguide path length from
the input port to each radiating slot is the same; thus no change in beam position
with frequency will occur. Another advantage is that all radiating slots experience
the same amount of waveguide dispersion. The radiating waveguides should be
along the squint angle so that the aperture is broadside. This specifies the angle
between feed and radiating guides.

9.2.2.5 Frequency scanning: Exactly opposite to the previous Section where


beam scanning is removed through use of equal line lengths is frequency scanning
wherein the interelement path length is increased. This increases the scan produced
by a given frequency change. Although frequency scanning is most often used in
planar arrays, the principles will be discussed here. Assuming that the radiating ele-
ments spaced d apart are connected by a serpentine (snake) feed of length s and
wave number 0, the wave front is defined by:
fa*sin0m = $s-2mn (9.175)
See sketch in Fig. 9.67. To avoid an endfire beam, kd > /3s — 2rmi. More than one
main beam can be avoided (see Section 9.1) by keeping:

7 < , • \ a (9176>
A 1 + sin 6m
Linear arrays 97

The phase equation is often written as:


s X mX
(9.177)
dXg d

radiating elements

Fig. 9.66 Equal-path-length feeds


a Symmetric
b Asymmetric

Clearly larger s/X, and correspondingly larger m, gives a faster change of beam angle
with frequency. If the frequency scan passes through broadside there will be a sig-
nificant change in impedance due to the addition of all the element conductances.
Element conductances, or couplings, are determined just as for a TW array. If the
waveguide bends needed to make a snake are not well matched over the frequency
band, it may be necessary to calculate the input admittance using the techniques
developed for resonant arrays. The coupler reflections tend to produce a sidelobe at
the conjugate beam direction; the level of this reflection sidelobe depends upon the
coupling conductances, and is higher when the scan is closer to broadside.
In a straight TW feed losses are usually not important. But the extra length of
98 Linear arrays
the sinuous feed often introduces non-trivial loss. Maximum gain occurs when
as = 1, where s is the total snake feed length (Begovich13). The corresponding feed
loss is 4.3 dB. Maximum radar target tracking accuracy, a combination of gain and
beamwidth, occurs for as = 2, with 8.6 dB feed loss (Skolnik139).

radiating
slots

load
waveguide '
Fig. 9.67 Snake feed geometry

Dispersion in the snake feed will distort short pulses if the pulse length is less
than, or comparable to, the feed travel time. Detailed calculations have been made
by Bailin,10 and Tseng and Cheng.164
As an example, assume each serpentine loop to be five free-space wavelengths,
s = 5X0, and a waveguide wave number ratio of p/k = 0.6 at centre frequency, and a
slot spacing of half free-space wavelength. The beam angle is given by:
= 10[Vl-0.8/o//)2-0.6/o//] (9.178)
This is shown in Fig. 9.68. A 10% bandwidth covers a 117° scan.
Fig. 9.69 shows a typical folded waveguide snake feed, while Fig. 9.70 depicts a
linear feed used in a cylindrical parabolic reflector. Another way of packaging the
snake is in a helix. Here a helical waveguide with the outside wall open is machined
from a solid block; a cover is later fastened around the open top (Croney and
Foster32). Higher dispersion, and hence a larger slope of degrees per MHz, can be
obtained from a dispersive, periodically loaded waveguide, analogous to the wave-
guide delay lines used in travelling-wave tubes (Hockham and Wolfson80).

9.2.3 Corporate line and distributed feeds

9.2.3.1 Corporate feeds: Corporate feeds are named after the structure of
organisation charts: a single port divides into two or more ports, each of which then
divides into two or more, etc., until the desired number of ports (equal to the num-
ber of array elements) is reached. Most often each port is split into two ports; for
this type of divider the number of elements must be a power of two. Bifurcated T
waveguide or coaxial-line T junctions, or hybrid junctions can be used. The former
are simpler, but hybrid junctions reduce the effects of element impedance mis-
match. Typical hybrids are the short-slot hybrid (Riblet;128 Levy102) which uses two
Linear arrays 99
90-|

0--

-90
0.95 1.0 1.05
f/fO

Fig. 9.68 Frequency-scanning example


s = 5\, j3A = 0.6, d/KQ - 0.5

Fig. 9.69 Waveguide serpentine feed {courtesy: Hammer)


100 Linear arrays

guides with a contiguous narrow wall, the top-wall hybrid (Hadge64), the magic tee
(Young,185), and the ring hybrid (Tyrell165). Power dividers and hybrids can also be
implemented in stripline or microstrip (Shimizu and Jones;138 Levy;102 Young186).

separation
of parabolic
cylinder common horn over slots to
direct power from feed into
parabolic reflector

focal line of left half of


parabolic cylinder

-focal lines
-line source sinuous
feed array

Fig. 9.70 Frequency scanner: Linear array feed a parabolic cylinder {courtesy: Begovich)

power
dividers

Fig. 9.71 Corporate feed


Numbers indicate units of phase shift

A broadband TEM line coupler is that developed by Schiffman.136 Fig. 9.71


sketches a typical linear corporate feed using power dividers. Since the path from
the feed port to each element is the same, the corporate feed is broadband. Of
course, if hybrids with a 90° output difference phase are used, the bandwidth will
Linear arrays 101

be limited. Scanning is produced by inserting a phaser at each element as shown in


the Figure. Although each element is shown with a phaser, only JV — 1 are actually
needed since the absolute radiated phase is usually of no interest. To make a cor-
porate feed scan over a broadband, it is necessary to use variable line lengths instead
of phasers*. Since these are expensive (lossy and bulky) a compromise can some-
times be used, where the array is divided into contiguous sub-arrays. Each sub-array
is fed through time delay (line length) while elements in each sub-array are control-
led by phasers (see Fig. 9.72).

elements

phasers

time delays

Fig. 9.72 Hybrid time-phase corporate feed

9,2.3.2 Line Feeds: Linear feeds are simpler than corporate feeds, and phasers
for beam scanning can be placed in either the branch lines or in the main line. Fig.
9.73 shows these configurations. In the branch-line phaser connection the advan-
tage is that each phaser handles only 1/7V of the transmitted power. However, at
any scan angle, the basic interelement phase shift <i> = kd sin6 must be multiplied
by the number of the element. Thus left to right the phasers are set to modulo 2n
values of 0, 2$, 30 etc. With digital phasers and a microprocessor control this is not
a disadvantage. Or an adder chain can be set up to add the phase increment from
each phaser to the next. For the main-line phaser connection the advantage is that
all phasers have the same setting of (f>. But the first phaser handles essentially all the
power and the losses are in series. The last elements have a reduced excitation then,
unless the couplings are adjusted for cumulative power loss.
Mutual coupling between array elements changes with scan angle, producing im-
pedance mismatches except at the design angle. These mismatches may upset the
operation of the feed network, and in some cases may introduce systematic errors
which produce a set of sidelobes of predictable location and height. For example,
take an array with phasers in the branch lines (see Fig. 9.73a) and consider the
array to be transmitting. Some energy will be reflected from the element due to the
impedance mismatch with scan; this energy goes back through the phaser. Each

* Phasers as used here are modulo 2tr


102 Linear arrays
branch line is lightly coupled to the main feed line, so most of the energy is reflec-
ted at the main-branch coupling back through the phaser again and on to the ele-
ment. Thus, if the initial element excitation has phase 6, the first re-reflected wave
has phase 3$, the second re-reflected wave has phase 5$, etc. Each wave has lower
amplitude than the previous wave, and each represents a beam at a different direc-
tion in space. Thus these effects are systematic rather than random. Examples are
given by Kurtz and Elliot." This discussion assumes reciprocal phasers; non-
reciprocal phasers will also produce systematic effects that depend upon the phaser
reverse phase.

elements

phasers
v
in o—
main feed line

V V V V

Fig. 9.73 Linear scanning feeds


a Branch-line phasers
b Main-line phasers

Similar effects can be observed with phasers in the main feed line (see Fig.
9.73Z>), where element mutual impedance changes with scan are reflected into the
main line (Spradley and Odium143). For a detailed discussion of this work, see
Kummer.98

9.2.3.3 Distributed arrays: A distributed array is one where each element is


connected to its own receiver/transmitter unit. Frequently the unit consists of the
necessary filters, diplexer and/or circulator, preamplifier, mixer, IF driver ampli-
fier, final transmitter power amplifier and drivers, and phaser. With the antenna it is
called a module; the circuit and element are usually packaged together. These types
of modules most often find use in large two-dimensional arrays (see Chapter 10). A
modification on the distributed array connects the receiver/transmitter circuitry to
a sub-array. Distributed arrays have two attractive features. First, the use of many
power amplifiers may allow solid-state components to be used in place of a single
high-power tube. Secondly, graceful degradation occurs. If one module fails, the
gain and sidelobes change only slightly. Thus the mean time before failure of the
array (for a specified gain degradation) can be much larger than the mean time
before failure of a module (Hansen68).
Linear arrays 103
9.2.3.4 Array filling time: When a corporate feed, or an equal-path-length
travelling-wave feed, is used with short pulses, all elements are excited simul-
taneously. A series feed in contrast excites one element first, with the last element
excited later by a time equal to the feed travel time. If the pulse is sufficiently short
the array elements are never all excited simultaneously, thereby precluding forma-
tion of the usual pencil-beam pattern with high gain. Bailin10 has analysed this
situation, under the assumptions of no feeder dispersion, no higher modes, slot
admittance independent of frequency, no multiple reflections and no mutual coupl-
ing. In spite of this highly idealised model, the results are quite useful. Fig. 9.74
shows a pattern buildup due to array filling time for a 10-element, half-wave spaced
array. The pulse-length parameter is a multiple of the travel time between adjacent
slots. It may be seen that a pulse length of four travel times has roughly 20 dB of
gain loss. A reasonable approximation to the desired pattern, and gain loss of
roughly ldB or less requires a pulse length twice the overall array travel time.

0 10

0-010-

0001
30 60 90 120 150 180
angle from broadside,deg

Fig. 9.74 Pattern buildup for 10-element array {courtesy: Bailin)


steady state
f0 = Time for pulse front to travel between two adjacent slots

A scanned array experiences similar problems even when path lengths are
adjusted so that the path length from the feed port through each element to a wave
normal are equal. The entire aperture is not excited until the pulse has reached the
104 Linear arrays

furthest element of the array. A simple but powerful analysis of antenna transients
has been given by Tseng and Cheng.164

9.2.4 Phaser and sub-array lobes

9.2.4.1 Sub-array lobes: When an array of sub-arrays is scanned, the aperture


phase consists of a stair step curve where each step represents the phase over one
sub-array. The desired aperture phase curve, of course, is a linear curve of phase vs.
distance, with the slope of the line proportional to the beam position, sin0o. Thus,
a phase-error curve of sawtooth shape is produced (see Fig. 9.75). The tooth width
is the sub-array width W and the amplitude each side of the axis is /3, where |3 =
(kW/2)sin90. This error phase produces one or more salient lobes called quantis-
ation lobes (QL); the amplitude and location of these lobes can be found by inte-
grating the aperture error function. For large arrays the illumination may be replaced
by a continuous distribution for the purposes of calculation. Let the array have N
equal sub-arrays, with uniform excitation. Integrating over the TV steps:

exp[jmp+jp(2n-l-N)-jWp]dp (9.179)

where u = (Z,/A)(sin0 o - sin0), NX = (2n -N)/N, and N2 = (2/i - 2 ~N)/N. Now


call:
W W
v = — sin0, w = —(sin0o—sin0) = vo-v
A A.

Since L =NW, u = Nw, and |3 = nv0. The integral and summation can both be per-
formed in closed form yielding:

F(u) = - — sincTrz; (9.180)


N sin nw
The first factor is recognised in the pattern of a uniform array of N isotropic ele-
ments spaced W wavelengths apart. For W larger than one wavelength grating
lobe(s) will exist. These grating lobes have equal amplitude of unity and occur for
vv = m, m = 0, 1, 2 , . . . . The first of these is the main beam, for vv = 0. The
second factor is the pattern of a uniform line source of length W wavelengths, with
main beam peak at v = 0 = 0. Thus this sine beam weights the main beam and side-
lobes of the first factor. Since the grating lobes are spaced by unity in vv (or v), and
since the sine null-to-null beamwidth is two (in z>), the first grating lobe is always at
the edge of or in the sine main beam. For example, if the first factor main beam is
at 0O = 0 (no scan), the first grating lobe is at the sine-factor first null. As the main
beam scans it moves away from the sine beam peak, and the grating lobe moves into
the sine beam peak from the null. The level of the first quantisation lobe is closely
just that of the sine factor at the first grating lobe position: v = v0 — 1 (from the
definition of vv) and wgl = 1. The next quantisation lobe will occur at the second
grating lobe, v = v0 — 2, and the sine-pattern sidelobes there will give the QL level.
Linear arrays 105
Since the QL peaks are produced by the product of the first-factor grating lobe
and the slope of the sine factor, the exact QL position can be found by differenti-
ating eqn. 9.180 and setting the result to zero. However, a numerical solution
would be required, given N and Wfk. As long as N> 1 the GL beamwidth will be
small compared with the sine-factor beamwidth, and eqn. 9.181 will be accurate.
For a smaller number of sub-arrays, the exact beam peaks should be determined;
from these, exact QL levels can be calculated.

\KK|
1 N 1 \ \| \J \l
\

0
/

desired phase /
b

Fig. 9.75 Quantisation phase errors


a Phase error
b Stairstep phase approximation

The three important levels are then:


main beam: Fo = sine [irv0]
c J. /^T r- r , «M sin7rz; 0
first QL: Fql = sine [TT(V0 - 1)] = — — ~ =

second QL: F , , = sine [ir(7; 0 - 2)] = - ^ ^ - = - ^


n(vo-~2) (9.181)
Clearly if the scan angle makes v0 = | , the main beam and first QL are equal! This
gives an absolute upper limit on W/X and 0o:
106 Linear arrays

W
— sin 60 < \
A
Locations of the QL are closely at the grating-lobe locations:
ink
sin0 Q j = s i n 0 o - — > m=l,2,... (9.182)
W
The second QL is always smaller than the first and has a peak value of 0.2172 at
j3 = 1.7989. For small j3, the QL amplitude is approximately 0/TT. The QL amplitude
is with respect to the main beam amplitude at broadside. At a scan angle, since the
main beam is reduced with the effective aperture, the relative QL increases. Calcu-
lations have shown that tapered (Taylor) distributions give essentially the previous
result as long as N > 20.* Values of J3/TT for various QL levels are given in Table
9.14; since j5/n = (W/X) sin $0, various combinations of W/X and 60 can be used.
For example, a QL no higher than — 15dB (with respect to main beam broad-
side) can be produced by a sub-array width of 0.9X for ± 10° scan. Suppression of
the QL occurs when the sub-array widths are slightly randomised.

Table 9.14 Sub-array quantization lobe


PI* QL Main lobe
dB dB
0.0307 -30 -0.01
0.0535 -25 -0.04
0.0921 -20 -0.12
0.1562 -15 -0.35
0.1872 -13.26 -0.51
0.2620 -10 -1.00
0.4430 -5 -3.01
0.5 -3.92 -3.92

9.2.4.2 Digital phasers: A digital phaser allows the desired phase to be approxi-
mated within the least count of the phaser. With M bits the least count is 2n/2M.
Thus the use of digital phasers for scanning also approximates to the ideal linear
phase over an aperture by stair steps, and the error curve is a sawtooth. However,
only the discrete points corresponding to the array element positions are important.
Two cases can be distinguished. If the element spacing and scan angle are such that
the number of elements is at least twice the number of steps, each step will have
two elements or more, and the discrete case approximates to the continuous saw-
tooth. For this case a discrete QL will be produced as in the sub-array situation.
The other case is where the number of elements is less than the number of steps,
which gives the discrete phase distribution a random character. First the oversampled

* Courtesy of Mr. C. L. Gray of Goodyear Aerospace


Linear arrays 107

case will be discussed. The end-to-end phase of an JV-element array is


(N — l)kd sin d0. Let the number of elements per phase step be H:
N
H (9 183)
'
If the element spacing and number of bits are given, the range of scan angles that
allows i / > 2 can be found. For most situations the angles are small. When H is
between 1 and 2 there is a transition between the random phase error with raised
sidelobes and the correlated phase error producing a quantisation lobe. The tran-
sition is only complete when H > 1. Step width w, which governs the QL angle
Bq, is given by:
(9 i84)
^ U
The peak sawtooth phase error j3 is:
-
0 = ~M (9-185)

The QL amplitude result obtained previously can now be used to find QL in terms
of $. Table 9.15 gives the results and it can be seen that 4 to 5 bits are needed for
low sidelobes, but 3 bits are often adequate. The QL amplitude can be significantly
reduced if a small random phase is added to each phaser. The QL is then de-
correlated and the overall sidelobe level is raised.

Table 9.15 Phaser quantisation lobe


Phaser bits QL
dB
2 -10.5
3 -17.1
4 -23.6
5 -29.8
6 -36.0

Case two, with random errors, can be evaluated by using an approximate vari-
ance for the phase of |$ 2 :
2
o2 = — (9.186)
3(4 M )
The RMS sidelobe level for uniform illumination referred to broadside is o2/N (see
Section 9.1.5). At a scan angle #0, the main beam energy is reduced approximately
eos2#0; so the RMS sidelobe level with respect to the main beam is increased. Fig.
9.76 shows the RMS sidelobe level referred to broadside. Use of an aperture taper
will also increase the sidelobe level. Along with the increased sidelobes due to
108 Linear arrays

random phase quantisation goes a modest gain loss:

(9.187)

Table 9.16 gives this loss for various numbers of phaser bits. From this Table, three
or more bits should be adequate.

phaser bits = 2
-20-

dB

-30-

-60
too
number of elements

Fig. 9.76 RMS sidelobe level due to random phaser quantisation error
Uniform array at broadside

Table 9.16 Phaser gain loss


M Random loss Regular loss
dB dB
2 1.000 0.912
3 0.229 0.224
4 0.056 0.056
5 0.014 0.014

The quantisation lobe reduces the main beam gain, and this gain can be shown
(Allen et al.4) to be approximately:

^ sine (9.188)

Values are given in Table 9.16 where it can be seen that 3 bits are adequate for
most arrays.
Linear arrays 109

9.3 Array elements

Arrays may be composed of low-gain elements such as slots or dipoles, or of


moderate-gain elements such as reflectors, horns, spirals, log periodics etc. The
beamwidth of the latter imposes restrictions on the array scanning range, and also
on the phasing due to quantisation lobes as shown in Section 9.2.5. Since dipoles
and slots are most often used in arrays, the relevant features are included in this
Section. Also discussed are multi-mode elements, designed to compensate for
mutual-impedance changes with scan. Printed-circuit elements, using either strip-
line or microstrip, have become important for many narrow-band applications.
Basic principles of these are covered.
Mutual coupling between array elements is discussed in detail in Chapter 10.

9.3.1 Dipoles

Because of its simple construction from two rods or tubes with a feed in between,
or two printed-circuit strips, the thin dipole is popular when the bandwidth is
roughly 10-20% or less. A strip dipole, where strip thickness is small compared to
strip width, is equivalent to a cylindrical dipole of radius a equal to one-fourth the
strip width w (Lo103). The effects of a dielectric sheet used as a substrate will be
discussed later. A half-wave dipole has a pattern symmetric about the dipole axis,
and with Q measured from the axis it is:

sin 6
The half-power beamwidth is 78.1°, and the directivity is 1.64 = 2.15 dB. As the
dipole length shortens, the pattern approaches that of a short dipole, sin#, with
half-power beamwidth of 90° and directivity of 1.5 = 1.76 dB. For lengths longer
than half wavelength, the pattern sharpens, then breaks up. At / = 0.625 X, the main
lobe is broadside, and the two sidelobes are small. But at / = 0.75X, the sidelobes
(at 45°) are larger than the main beam. Finally the full-wave dipole with sinusoidal
current distribution has a null at broadside. Pattern bandwidth then goes from a
short dipole (limited by impedance) to roughly 0.6X length.
Impedance bandwidth is limiting for thin dipoles. It was shown by Abraham in
1898 that a vanishingly thin dipole has a sinusoidal current distribution. The Carter
zero-order impedance theory discussed in Section 9.4 is adequate for mutual-
impedance calculations, and for self impedance of very thin (//a> 1000) dipoles.
Full-wave dipoles according to this theory have zero feed current, so a higher order
theory is needed for them. However, owing to their high input impedance and multi-
lobe pattern, full-wave dipoles are rarely used. For antennas in the vicinity of half
wave in length, the King second-order theory (King94) agrees well with experiments
as do the moment-method calculations (Harrington75). Unfortunately, no satisfactory
simple function fit to the impedance vs. frequency has been possible. The designer
can start with curves such as Fig. 9.77 from King,94 but should finish with
110 Linear arrays

6000-1
a = 2 In (I/a)
vQ = 20

Fig 9.77 Second-order resistance of cylindrical antenna king-Middleton expansion


a Dipo/e resistance (after King)

measurements or moment-method calculations. Owing to the 'capacitance' of a


cylindrical conductor, resonance occurs for lengths below half wave. Fig. 9.78
shows the shortening factor vs. X/a using zero-order theory. The graph stops at a
radius of X/100 as the zero-order theory is not accurate for fat dipoles. For these
cases the King data referred to earlier can be used, but a careful measurement pro-
gram is better. With the advent of network analysers, antenna impedance measure-
ments have been greatly simplified.
Most arrays work around broadside, for which dipoles parallel to the array face
are appropriate. Endfire or near-endfire arrays would typically use dipoles normal
to the array face. Here monopoles over a ground plane could be used; monopole im-
pedance is half that of the equivalent dipole, but the pattern is the same. Parallel di-
poles are often fed by a balun, which converts the symmetric twin-line dipole feed
to the coaxial line usually used behind the array face. Cylindrical dipoles are often
spaced A/4 or less above a ground plane, with a balun normal to the ground plane
supporting the dipole. The coax feed line then penetrates the ground plane. Either
Linear arrays 111
6OOO-1

0 1 2 3 4 5 6 7

(b) Poh
Fig. 9.77 Second-order resistance of cylindrical antenna king-Middleton expansion
b Dipole reactance (after King)

types I or III baluns* can be used. Circular or adjustable polarisation can be


provided through the use of crossed dipoles connecting to four balun posts with
two coax penetrations. These are then phased for the desired polarisation. A simple
way of producing circular polarisation uses crossed dipoles connected to a single
balun. One dipole is larger and thinner, the other shorter and fatter, so that the res-
pective phases are ±45°. A dipole over a ground plane is equivalent to the dipole
and its image (at a distance twice that of the dipole over the plane); thus this 'two
element array' provides directivity. The pattern for half-wave (image) spacing,
6 measured from the array axis (normal to the ground plane) but not including the
dipole pattern, is sin(kh cosO). As the spacing decreases the pattern approaches
cos0. Directivity increases slowly as h (height above ground plane) decreases. How-
ever, since the image current is reversed, as h decreases, the heat loss increases. Thus
the gain actually peaks. Fig. 9.79 shows directivity vs hfk, and also gain for several
loss factors. The loss resistance shown is the equivalent resistance at the terminals.
This is related to the surface resistance by multiplying the latter by the metallic
* As defined circa World War 2 by Reich. 122
112 Linear arrays

0-98-

0-96-

0-9/.

1
0-92-

100 103

Fig. 9.78 Dipole shortening due to diameter

10 n

0.5 0.10 0.15 0.20 0.25

Fig. 9.79 Effect of loss on dipole over screen


Linear arrays 113

path length and dividing by the path width. For sinusoidal current the path length
is A/4 while the width is the dipole circumference or width:

RL = te = M (9.190)
Sna Aw
For example, a printed circuit dipole at 6000 Mhz made of aluminum, and with a
length/width ratio of 62.5 has a terminal resistance of 0.50O. There is then a
spacing between dipole and ground screen that maximises the gain.
Direct (metallic) coupling of a dipole to a transmission line is not necessary;
electromagnetic coupling may be used (Forbes55). Here the shorted dipole is placed
above a two-wire feed line at an angle to the line.
Dipole bandwidth increases as the dipole becomes 'fatter'; for very fat configur-
ations a bow-tie shape has demonstrated excellent bandwidth. Only numerical
methods such as the moment method have proven useful in analysing such fat di-
poles.
Figs. 9.80 give impedance design data for bow-tie monopoles vs. electrical half
length in degrees as calculated by Butler et al2S Much broader bandwidth can be
achieved by utilising a three-dimensional structure, the open sleeve dipole (King and
Wong93). The sleeve is open as it does not surround the dipole, but consists only of
two tubes on opposite sides of the dipole. Fig. 9.81 sketches a linearly polarised
open sleeve dipole. Crossed dipoles with crossed sleeves for CP can also be construc-
ted. The crossed sleeves on each side can be replaced by a metallic circular disc.
Bandwidth, for VSWR<2 can be nearly an octave. These dipoles have been
arrayed over a ground plane with excellent results.

9.3.2 Waveguide slots

One of the most widely used radiators for arrays is the waveguide slot, simply a nar-
row slot cut into the broad wall or edge of a rectangular waveguide. Slots can be
accurately milled, especially with numerically controlled machines, and the wave-
guide provides a linear feed that is low loss. The precise control of aperture distri-
bution afforded by a slot array has led to their replacing reflector antennas in many
missile and aircraft radar systems. A narrow slot along the centre line of the
waveguide broad wall does not radiate. To produce radiation the slot must be dis-
placed (toward the edge) or rotated (about the centreline). Similarly, a slot in the
narrow wall does not radiate if it is normal to the edge; rotating the slot couples it
to the waveguide mode. Since an edge slot must usually be 'wrapped around' to get
resonant length, displacing such a slot is impractical. Fig. 9.82 shows the three most
important types. The displaced broad wall slot is often called a shunt slot as its
equivalent circuit is a shunt admittance across the feed line. The rotated series slot,
centred on the broad wall, is a series slot, as its equivalent circuit has a series im-
pedance. And the edge slot is a shunt-type slot. The edge slot and displaced broad-
wall slot are most often used, with polarisation often the deciding factor. In an edge
slot the E polarisation is along the guide axis, while for displaced slots it is across
114 Linear arrays

the axis. Pattern behaviour of waveguide slots is close to that of slots in an infinite
ground plane, except at angles near grazing where edge effects are important. How-
ever, in a linear array of waveguide slots spaced in the rough range of 0.5 to 0.8

600- 80

500-

400-

300-

200-

100-

0
0 30 60 90 120 150 180 210 240 270 300 330 360
A,deg

300-

-300-
J
A.deg

Fig. 9.80 Bow-tie monopole impedance (after Butler et al)

free-space wavelength, the slot phases must usually be alternated to correct for the
waveguide phase. For example, in a resonant array with elements spaced X^/2 apart,
the elements will alternate phase unless every other element has coupling reversed.
Linear arrays 115

Thus such arrays have slots alternating on each side of the broad-wall centreline; or
edge slot angles alternating from positive to negative. Now the array pattern is com-
posed of an array factor with double spacing, and an element consisting of a pair of
slots. For longitudinal slots the lobes can be controlled through element spacing.
With edge slots, however, cross-polarised lobes are produced at certain angles. These
lobes can be reduced by replacing each slot by a closely spaced pair of slots with
smaller inclination angle.

0-U1 in.SEMIRIGID
COAXIAL CABLE
REFLECTOR
SURFACE

f COAXIAL
INPUT
Fig. 9.81 Open sleeve dipole {after King and Wong)

longitudinal
displaced

inclined

edge

Fig. 9.82 Waveguide slots

Admittance (or impedance) of a waveguide slot consists of an external contri-


bution which can be computed from the equivalent slot in ground plane or Babinet
equivalent dipole, and an internal contribution due to energy storage in evanescent
modes in the guide around the slot. Broad-wall slots are simpler and will be discus-
sed first. Since the internal contribution is reactive, the conductance of a resonant
slot can be found approximately from the Babinet dipole and from the coupling to
116 Linear arrays

the TEOi mode. Stevenson149 developed the formula:

- = —~^-sin2—cos2— (9.191)

where Ro is the dipole resistance and x is the slot offset. This formula resulted from
the assumption of a cosine field distribution in the slot. Das and others33 observed
that a sinusoidal distribution is more appropriate for a slot, which yields the
modified Stevenson formula:
=
v ^-sin2-tcos^/-cosM]2 (9.192)

The unmodified formula is less accurate when the slot length / is not close to A/2.
Variational formulation of the susceptance problem was made by Oliner.114'11S He
obtained closed-form simplifications which are of use, but not sufficiently accurate
for array design. With the availability of powerful computers, a more complete
evaluation of variational forms can be made. Yee184 developed such a formulation.
A different type of solution was developed by Khac,90'91 who wrote coupled
integral equations representing external and internal electric fields. The coupled
integral equations are then solved by the moment method, using pulse-expansion
functions and delta-testing functions. In both, wall thickness is included via Oliner's
method of coupling external and internal fields by a waveguide transmission line,
where the waveguide cross section is the slot and the waveguide length is the wall
thickness. Satisfactory agreement with measured data has been realised with both
approaches.
To understand the behaviour of slots, and to put slot data in a format useful for
array design, the 'universal' curves can be utilised. Longitudinal broad-wall slots are
again used as they can be more accurately calculated. Using the Yee variational
method, corrected and improved by Goodyear Aerospace Corp.,* admittance of
slots in WR-90 waveguide was calculated. First, resonant slot length vs. slot offset
and resonant conductance are calculated. Offset x is conveniently normalised to
guide width a, while resonant length lr is normalized to free-space wavelength Xo.
Fig. 9.83 shows curves fora = 0.9, wall thickness = 0.05 and slot width = 1/16, all
in inches. Frequencies of 9.375 and lOGhz are shown, with b = 0.4 and 0.2. Stand-
ard WR-90 has 0.4 height, but the half-height guide is of considerable interest for
receiving arrays. Resonant conductance is shown in Fig. 9.84 for only one case; it
matches very well the Stevenson formula given earlier.
Next the calculated admittance is normalised by resonant conductance Gr with
slot length normalised by resonant length. The result is the 'universal' curves
developed by Kaminow and Stegen,88 (see Figs. 9.85 and 9.86). The curves are uni-
versal in the following sense. To first order the variation of Y/Gr with l/lr and with
x/a is independent of 0/fc, that is of frequency. Thus measurement or calculation of
data for a given waveguide size at one frequency is adequate. For small offsets (0.05

Courtesy Gary Brunner


Linear arrays 117

and below) the normalised admittance variation is also independent of a/b. How-
ever, for larger offsets data should be obtained for the exact a/b. The variational
calculation of admittance is much too slow to use with iterative methods such as
those described in Section 9.2; polynomial fits are more suitable. However, the uni-
versal curves are difficult to fit with one expression. Dividing up the curve allows a
more accurate fit, but care must be taken to avoid slope discontinuities at the joins
as these affect the gradient methods used in Section 9.2. An alternative method uses
impedance rather than admittance. When slot impedance is normalised to resonant
resistance, the curves are as shown in Figs. 9.87 and 9.88. Resistance is a straight
line (to first order), with the slope independent of a/b. There is a slope change with
fi/k as shown. The reactance curve is a slightly curved line, which again is independ-
ent of fi/k9 and for small offsets is independent of a/b also. For larger offsets the
reactance curves depend on bothx/a and a/b. Thus the impedance curves are just as
'universal' as the admittance curves, and are much easier to fit with polynomials.

05161
9375 MHz, o = 0 - 9 , b = 0 2

0508-

0-500-

0492-

9375 MHz, a = Q 9 , b = 0 4
0-484-
10 GHz. a = 0 9 , b = 0-4

0476-

0 468

0 460
0 005 010 015 020 0-25
x/a

Fig. 9.83 Resonant length of longitudinal slot

If data are not available on a particular size and wall thickness of waveguide to be
used in an array, it is necessary to either calculate or measure sufficient data to plot
universal curves. The impedance curves are preferable as they are slowly varying.
Edge slots are difficult to analyse due to their wrap-around nature. For the same
118 Linear arrays

0 1-

001

0001
005 010 015 0-20 0 25
x/a

F ig. 9.84 Resonant conductance of longitudinal slot


Stevenson
o Yee
Linear arrays 119

reason the wall thickness has a significant effect on admittance. For reduced-height
guide, where the wrap-around is severe, even the pattern is difficult to calculate. At
present, no theories for edge slots have appeared, using either variational methods
or moment methods. Array design is based on measurements, and, even here, edge
slots are difficult. Because of the strong mutual coupling between edge slots (as
compared with displaced broad wall slots) incremental conductance is usually
measured.

0.6

x/a=0.05

1.0

0.8

0.2

-0.6 "

0.90 0.92 0.94 0.96 0.98 1.00 1.02 1.04 1.06 1.08 1.10

Fig. 9.85 Longitudinal slot admittance


a = 0.9, b = 0.4, 9375 Mhz
Stegen o x/a - 0.07
a 0.17

That is, a series of resonant slots is measured; then one slot is taped up and the
remaining slots are measured again. The resulting incremental conductance
(Watson173) is that of a resonant slot in the presence of mutual coupling. Trial and
error is necessary to find the resonant length for a set of slots for a given angle,
however. For inclination angles below 15°, the resonant conductance developed by
Stevenson (Watson173) varies as sin2:
4S0a/b
(9.193)
7T$/kR0(2a/X)4
This is shown in Fig. 9.89, the error in this curve for larger angles is not important,
as incremental conductance should be used. It appropriately follows a sin20
behaviour, especially for small angles, with the incremental value larger than the
120 Linear arrays

0.6 -

*^\\x/a=0.05
0.4 0.2 \ X

0.2
/A V
// \ - QA

0 -
// \ \ \x/a = 0.2

V \v
0.05 \v \v - O.I
02 / /0.05 \
-0.2 /

- 0.

-a*
• • ' ' — " ~~
- o.
-0.6
0 i i i i l 1 ( 1 1

0.90 0.92 0.94 0.96 0.98 1.00 1.02 1.04 1.06 1.08 1.10

Fig. 9.86 Longitudinal slot admittance


a = 0.9,/? = 0.2, 9375 Mhz

12-

0-8-

(U-

-0-4-

-08-

-092

0-90 092 0-94 096 0-98 100 102 1-04 106 108 1 10

Fig. 9.87 Longitudinal slot impedance


3 = 0.9,6 = 0.4
Linear arrays 121

single slot conductance. Resonant lengths will be different in an array, of course; so


extensive measurements are necessary to develop sufficient data for design.
Although many industrial organisations have done this, almost no measured slot
data have been published.

Fig. 9.88 Longitudinal slot impedance


a = 0.9, b = 0.2, 9375 Mhz
x/a = 0.05

The wrap-around nature of edge slots can be avoided by using an H slot (Chignell
and Roberts29), where a slot normal to the guide edge is augmented by slots at each
end that are parallel to the edges. Thus the slot looks like an H. By making the out-
side arms asymmetric the coupling can be varied, and in fact circular polarisation
can be achieved (Hill79).
Other types of slots such as probe coupled, iris excited, and crossed are discussed
by Oliner and Malech.115

9.3.3 Scan compensated elements

It was observed by Wheeler175 that a current sheet exhibits a radiation resistance


that increases with if-plane scan as sec0 and decreases with E-plane scan as cos0.
The current sheet, composed of short dipoles each with pattern cos0 in theF-plane
and unity in the /f-plane, causes this behaviour. An 'ideal' element pattern was indi-
cated by Wheeler to be cos1/20, where 0 is a conical angle from broadside. This is a
type of Huygen's source; an approximation is given by a superposition of a dipole
(electric) and a loop (magnetic dipole) which gives a pattern:
122 Linear arrays

f(6) = K l + cos0) = cos 2 - (9.194)

For a linear array the pattern is not a pencil beam, and thus the element orientation

0-1 -

001
10 15 20 30
inclination angle, deg

Fig. 9.89 Resonant conductance o f single edge slot


WR-90: 9375 MHz; slot width: 1/16 in

is important. From the grating lobe series (see Chapter 10) the active element
pattern can be written as a series containing ideal element patterns evaluated at the
grating lobes. When this is applied to a linear array of short dipoles (current line),
Linear arrays 123
171
the following ideal element patterns result (Wasylkiwsky and Kahn ):
collinear F(d,(p) = 1
(9.195)
parallel /f-plane F(0,0) = 1/(2 - cos20)
Little has been done on approximating the parallel //-plane pattern. However it is
possible to better approximate the planar array ideal cosi/2# pattern by using TE
and TM waveguide modes instead of electric and magnetic dipoles. This multi-mode
element is simply an open-ended rectangular or round waveguide arranged to sup-
port the proper mix of TE and TM modes. Such elements are able to greatly reduce
mutual-coupling impedance changes with scan.
Compensation of active impedance changes with scan angle can be done several
ways. The three most important methods are baffles and interconnections in the
feed network (see Chapter 10) and multiple mode elements. Knittel95 has surveyed
wide-angle impedance matching (WAIM) techniques. One way of looking at the
multiple mode element is that the combination of modes approximates the ideal
element pattern. Another viewpoint is that one group of modes has a mutual coupl-
ing scan behaviour opposite to that of another group of modes. The proper mix of
modes is produced by one or more dielectric sheets placed in front of the array
face and a dielectric plug in each element. The external dielectric sheet for scan
matching was developed by Magill and Wheeler107 and analysed by Wu and
Galindo182 and by Lee and Mittra.101 Use of a dielectric plug with two modes for
wide-angle matching was developed by Tang and Wong156 and Galindo and Wu.60
The mix of modes that is needed depends upon the waveguide dimensions and on
the lattice parameters. A typical mode set might be TE 10 , TE 2 o,TE 3O ,TE 11 , TE01,
TM n . With a dielectric plug in the waveguide the TE2o mode will likely be pro-
pagating; all other higher modes will be evanescent. The plug should be half wave
long (in loaded guide wavelengths) at centre frequency to properly locate the active
admittance on the Smith chart. Then the guide dimensions are adjusted to give a
high susceptance for the TE2o mode looking into the matching network. Steps in
guide height can then be used for centre-angle impedance matching; additional
dielectric plugs can also be used for matching. Fig. 9.90 sketches a typical guide
configuration, including two sheets over the array face. The design process for
WAIM elements is based on work by Diamond,37 Knittel, Hessel and Oliner.96
Additional references are Amitay and Galindo;7 Stark, Tang and Wong;144 Diamond
and Knittel;38 Wong et al;179 Hansen;70'72 and Amitay, Galindo and Wu.6
The analysis uses the unit-cell concept pioneered by Oliner116 (see Chapter 10).
Since the unit cell at scan angles other than broadside has impedance walls, the
modes are USE and LSM rather than TE and TM. It is necessary to assume values of
all physical parameters to start the process: slab dielectric constants and thick-
nesses, plug dielectric constants and lengths, waveguide heights and widths and, of
course, the element lattice dimensions. A mode set for the waveguide element is
then assembled. Choosing a reasonably good set of all these parameters requires con-
siderable experience, gained mostly by running various cases and then examining
124 Linear arrays

the results. With all parameters set, the unit-cell admittance is calculated for each
waveguide mode, assuming unit amplitude of the latter. Many modes will be
required in the unit cell to match the waveguide mode field at the aperture (the
waveguide open end), and for each unit cell mode the admittance must be calcu-
lated. Starting with the value for large distance (from the slabs), this admittance is
transformed through the dielectric slabs, and then combined with the other unit
cell modes. Numbers of modes typically are 100 to 300. Longitudinal-section
modes must be converted to TE and TM modes so that mode admittances can be
matched. Next, starting with an incident TE10 mode in the waveguide, the aperture
admittance of each waveguide mode is calculated. Although the mode admittances
must be equal across the aperture, the mode amplitudes depend upon the total
admittance. Thus, a set of simultaneous equations, one for each waveguide mode,
can be solved. Typical results are shown in Fig. 9.91 which is for 0.403X0 square
waveguides in an equilateral triangular lattice (Tsandoulas and Knittel161). Band-
width for VSWR<3 is 25%, and the scan range is roughly a quarter hemisphere.
An interesting design principle resulting from many calculations by these authors
is that polarisation coupling (for a dual polarised array) is reduced as the element
size is increased. Thus when multi-mode elements are used they should be as large
as the lattice allows. For a single polarisation, bandwidth of greater than 50% with
VSWR < 4 and scan over nearly \ hemisphere has been obtained (Tsandoulas160'162).

waveguide
steps

Fig. 9.90 Typical matching section and unit cell

An iris can be used in each waveguide radiator, below the aperture, in addition
to dielectric plugs, as an aid to matching. Excellent results have been obtained by
Lee and Jones and by Lee.
Another example of the excelllent single-plane scan compensation results obtain-
able is in Fig. 9.92 (Tsandoulas162). This array uses open-ended waveguides in a rec-
tangular lattice for i?-plane scan. All dimensions have been optimised using a multi-
mode element code described earlier.
A newer technique can be used on single-mode elements. It uses a spatial filter
consisting of one or more dielectric slabs over the array face where the transmission
Linear arrays 125

10-

0.5.
\
\
N
x
(I \
-e-
c (I 1 sin9 cos<(>(H-plane)

IU endfire scan limit

/
/
\ /
y
0-5- / grating lobe

scan volume

10-

Fig. 9.91 VSWR = 3: Contours


Polarisation
E parallel to plane of scan
E perpendicular to plane of scan

15

1 U-

i13
CO

1-2-

1-1-

10
10 20 30
scan angle 8 >

Fig. 9.92 E-plane scan at high and low frequencies (courtesy: Tsandoulas)
bandwidth = 12%
a/\0 = 0.670
dy/\Q = 0.500
ib/\0 =0.420
\ 0
lb/\0 =0.410
126 Linear arrays
coefficient of the filter varies with angle and with polarisation in such a manner as
to modify the active element pattern toward cos B. See Chapter 10.

9.3.4 Printed-circuit antennas

Stripline slots: Antennas constructed in microstrip or stripline are printed-


circuit antennas, the principal advantages of which are low fabrication cost and ease
of integration with feed lines and components. Slots in stripline are typically
narrow rectangular resonant slots cut into the outer conducting plate of the strip
line, with the strip centre conductor oriented normal to the slot and offset from the
centre. Slot length is adjusted for resonance. Conductance can then be calculated
(Breithaupt20). For example, a metal-clad dielectric sheet would be etched to pro-
duce the slot radiators, then a double-clad dielectric sheet would be etched on one
side to produce the feed network. These two would be fastened together to make
the stripline slot array. Although the external (radiation) behaviour of a stripline
slot is similar to that of a slot in an infinite ground plane, the internal admittance is
strongly affected by the closely spaced ground plane underneath. The net result is
to severely restrict the impedance bandwidth.

Fig. 9.93 Boxed stripline array and centre strip (courtesy: Park)

Another stripline slot configuration uses boxed stripline (stripline with shorting
edges parallel to the strip conductor). Each array stick looks like waveguide (see
Fig. 9.93), except that the slots are all on the centreline. The strip conductor is
symmetrically angled below each slot, with the coupling proportional to the strip
inclination angle. Snaking the strip conductor with the slots along the centreline
eliminates unbalance and loading problems experienced in the earlier work of
Strumwasser* where the strip was straight and the slots angled. The slots here are

* Unpublished Hughes report circa 1952


Linear arrays 127

one wavelength apart (in the stripline medium), so no phase reversals are needed.
Mutual coupling is accommodated in the design by techniques described in Section
9.2. A 10-element array is shown in Fig. 9.93 (Park and Elliott118).

Microstrip patches and stubs: Microstrip antennas consist of dipoles or other


resonant metallic shapes in the top microstrip conductor. A microstrip dipole is
affected by the dielectric so that the resonant length and impedance are changed
from the free-space values. In principle, the current distribution along the dipole (or
the field along the slot) is a solution to the boundary-value problem, as yet
unsolved. A start has been made by Alexopoulis3 in developing a Green's function
for an infinitesimal dipole on a grounded dielectric slab, i.e. a short microstrip di-
pole. This is then used to calculate mutual impedance between two such dipoles, so
that a moment-method solution can be obtained. Here a microstrip dipole is repre-
sented by a set of overlapping short dipoles, with the simultaneous equations then
solved for the short dipole currents from which the impedance is immediately
determined. This also allows the shortening factor due to the dielectric to be deter-
mined. Since a dipole is only poorly represented by a transmission-line model, the
dipole shortening factor is not close to the microstrip velocity ratio. As in the case
of fat dipoles, resonant length should be determined by measurement. The earliest
reference to microstrip antennas appears to be that by Deschamps and Sichak.36
The patch shape has proven superior to the dipole, probably due to its smaller size.
It consists in its simplest form of a circular or square plate, fed at the edge by a
microstrip line. Alternatively, the patch can be fed by a coaxial cable connected to
the underside of the ground plane, with the centre conductor extending up to the
patch. Early rudimentary theories which calculated admittance from a model with
two radiating slots (the magnetic walls) connected by a low-characteristic-
impedance transmission line (the patch) have been superseded by modal theories.
Because microstrip is thin in wavelengths, the volume between the patch and the
ground plane acts like a TM resonator, where the edge dimensions are the appro-
priate length for the desired mode. The top and bottom walls of this 'cavity' are
metallic or electric walls; to a good approximation the side walls (between the
patch edge and the ground plane) are magnetic walls. Such a cavity supports a
dominant TM mode. The electric fields over the magnetic walls (between ground
plane and patch) act as virtual radiators; for a rectangular patch the field is large on
opposite walls so that the radiation pattern can be accurately calculated from this
model (Lo etal. ;106 see sketch in Fig. 9.94). Admittance computed from this model
gives a curve on the Smith chart that fits measured data well, except for a shift in
frequencies. This is because the magnetic walls are not perfect; there is a real com-
ponent of wall admittance due to radiation and an imaginary component due to the
discontinuity. Carver26 used the microstrip-discontinuity susceptance formula
developed by Hammerstad,66 with a series of cavity modes to match this admittance
boundary.
This made a considerable improvement in calculation of resonant frequency,
with typical accuracy of 0.5% for a rectangular patch and 0.1% for a circular patch.
128 Linear arrays

Resonant resistance values are predicted to typical accuracies of 10% and 3%, res-
pectively. Recently James86 has developed a variational calculation for the edge
admittance, and when this is incorporated into the patch modal solution the accur-
acy of the results should increase. Thus, one may conclude that microstrip antenna
design is now on a sound theoretical basis.

metal .

dielectric ? \ electric field

metaK

Fig. 9.94 Patch microstrip antenna


a Top view
b End view

Because of the close spacing between patch and ground plane, bandwidth of
microstrip antennas is limited. An accurate calculation of bandwidth includes loss
terms due to radiation, copper loss, dielectric loss and a very small term represent-
ing surface waves excited along the dielectric surface. Using the modal theory, the
radiation Q can be obtained from the TMm mode (Carver26). Then the dielectric
and copper loss Q$ can be obtained from the mode fields. Finally the Qs are com-
bined in parallel to give bandwidth: A / / / = l/Q. This has been done using the
magnetic-wall assumption at the gaps, with the result shown in Fig. 9.95 (Derneryd
and Lind34). It can be seen that bandwidth is nearly linearly proportional to micro-
strip thickness in guide wavelengths; a thickness of 0.05 gives about 4% bandwidth.
For very small thicknesses, the copper losses become large, which increases the
bandwidth! For practical purposes, however, the variation is linear. There is a sort
of fundamental limitation here similar to that of Chu: a given substrate thickness
yields a fixed bandwidth, essentially dependent only on thickness.
Bandwidth can be increased by the simple expedient of extending the patch
above the substrate. It is not desirable to increase the entire substrate thickness due
Linear arrays 129

000 0 15

Fig. 9.95 Bandwidth of microstrip antenna {courtesy: Derneryd & Lind, 1979)
VSWR < 2
tan (6) = 10" 3
a = 10 7
er = 2.52
calculated
• measured

metal

dielectric

b metal
Fig. 9.96 Two-layer microstrip for increased bandwidth {courtesy: Hall et al.)
a Top view
b Side view
130 Linear arrays

to direct radiation from the feed conductors. Fig. 9.96 from Hall, Wood and
Garrett65 shows a way of coupling to the raised patches. Bandwidth can be
increased by a factor of 2 or 3, but the low-cost fabrication and smooth array sur-
face have both been lost.
Dual orthogonal polarisation can be achieved simply by feeding a patch with two
properly phased feeds. However, circular polarisation can be produced by a single
feed. Feeding a circular or square patch at the edge produces linear polarisation. At
one time it was thought that a more complex shape, such as a five-sided patch,
might be needed to radiate circular polarisation. CP can be produced by a nearly
square patch fed at the corner, where one pair of sides (magnetic walls) resonates
at a slightly higher frequency than the other pair. If the phase difference at the
centre frequency between the pairs of sides is n/2 circular polarisation results (see
Fig. 9.97). Another scheme uses a square patch fed at the centre of one edge, with
an inclined slot cut out in the centre of the patch (Kerr89). Figs. 9.98 and 9.99
show the patch and the pattern measured with a rotating linear dipole. Truncating
opposite patch corners has the same effect in producing CP.

a+A

Fig. 9.97 Patches that radiate CP (courtesy: Kerr, 1979)

Feeds: Stripline or microstrip elements may be fed by a corporate feed, where


the power dividers or couplers use the stripline centre conductor or the microstrip
conductor. When the array is to be scanned there is usually not room for phasers
and power dividers and element coupling in one layer. In these cases a multi-layer
board is used. For example, one stripline might contain the corporate feed, with the
next layer containing phasers. Finally the top layer could be stripline slots or micro-
strip patches. Connections between boards are difficult; printed-through holes or
edge straps have been used. Fig. 9.100 shows a single-layer array of stripline slots
fed by a corporate feed. In this example the slots are fed by a centred quarter-wave
stub. The corporate feed uses line lengths rather than line widths for matching,
allowing loose tolerances.
Large arrays involve large corporate feed losses; series feeds offer an alternative.
Resonant feeds are limited owing to the combination of narrow-element bandwidth
and narrow-array bandwidth. When the required bandwidth is small this type of
array is excellent. Fig. 9.101 shows a 26-element centre-fed array with a 30dB
Taylor design.
Travelling-wave patch arrays may be used to provide an inexpensive array. Such
an array is designed by starting with a desired amplitude taper, then converting this
Linear arrays 131

Fig. 9.98 CP square microstrip radiator {courtesy: Kerr)

center-fed square radiator


rectangular polarizing
patch 1358MHz
I

Fig. 9.99 Typical axial-ratio pattern {courtesy: Kerr)


132 Linear arrays

to a set of patch conductances across the microstrip feed line. Using modal theory
each patch length (along the feeder) and width are chosen to give the required con-
ductance with the length then adjusted for match. Fig. 9.102 sketches a travelling
wave array built by Metzler; the actual array has 28 elements. The range of coupling
that can be achieved and the impedance properties of such series-fed patch arrays
need further investigation. Another travelling-wave configuration uses open-circuit
microstrip stubs as radiators, with stubs alternating left and right. The beam squint
can be removed without incurring a high resonant VSWR by using stubs in groups
of two with A/4 spacing. The 45° phase error increases sidelobes somewhat. Fig.
9.103 depicts such an array (James and Hall85). Stub width may be varied to
achieve an aperture taper.

Fig. 9.100 Strip line slot array corporate feed {courtesy: Waterman)

Fig. 9.101 26-eiement series-fed array for 1.415 GHz {courtesy: Metzler)

Integrated elements: At wavelengths of several millimeters to 0.1 mm the possi-


bility of constructing an antenna or an array element on the same substrate that
contains a detector (for receiving) or a source (for transmitting) becomes attractive.
Most of the work to date has used a substrate with detector. One configuration is a
travelling-wave vee antenna a few wavelengths long, with the vee angle chosen to
Linear arrays 133

maximise gain, as in a rhombic. The antenna and the coupling are all produced by
successive etches on the substrate. A second configuration uses a dielectric-rod
surface-wave antenna, with the rod consisting of a tapered-height rectangular bar
made on the substrate, and excited by a vee antenna. Another uses a reflecting
corner with a wire feed several wavelengths long, with the corner made as part of
the substrate. An array configuration uses a grid of parallel conducting strips with
an array of microbolometers between the strips. An overview of this embryonic
field is given by Schwarz and Rutledge.137

travelling wave feed

Fig. 9.102 Travelling-wave patch array {courtesy: Metzler)

fl
UU LT¥
Fig. 9.103 Travelling-wave array of microstrip stubs {courtesy of James and Hall)

9.3.5 Moderate gain elements


This class of array elements includes ail those with directivity/gain higher than that
of resonant dipoles and slots. Moderate-gain elements are of many diverse types,
including:
planar spirals
conical spirals
log-periodic
pyramidal log-periodic
horn including corrugated
scalar horn
Yagi-Uda
backfire
small paraboloid
ridged or TEM horn
helix including quadrifilar
rhombic
Beverage
These antennas have little in common with each other, nor are their array proper-
ties unique. Thus the reader is referred to other chapters that give considerable
detail on these moderate-gain antennas.
134 Linear arrays
9.4 References

1 ABRAMOWITZ, M., and STEGUN, L.: 'Handbook of mathematical functions' NBS, 1970
2 AGRAWAL, V. D., and LO, Y. T.: 'Mutual coupling in phased arrays of randomly spaced
antennas', IEEE Trans., AP-20, May 1972, pp. 288-295
3 ALEXOPOULOS N. G. et al: 'Radiation properties of microstrip dipoles', IEEE Trans.,
AP-27, Nov. 1979, pp. 853-858.
4 ALLEN J. L., et al: 'Phased array radar studies', Lincoln Lab. MIT. TR238, Aug. 1960
5 ALTMAN, J. L.: 'Microwave circuits', Van Nostrand Co., 1964
6 AMITAY, N. et al.: Theory and analysis of phased array antennas'. Wiley, 1972
7 AMITAY, N., and GALINDO, V.: 'The analysis of circular waveguide phased arrays, BSTJ,
47, 1968, pp. 1903-1931
8 ANDREASEN, M. G.: 'Linear arrays with variable interelement spacings', IEEE Trans.
AP-10, 1962, pp. 137-143
9 BACH H., and HANSEN, J. E.: 'Uniformly spaced arrays', in Collin and Zucker (eds.)
'Antenna theory Pt 1'. McGraw Hill, 1969, Chap. 5
10 BAILIN, L. L.: 'Fundamental limitations of long arrays'. Rept. TM33O Hughes Aircraft
Co., Culver City, California, 1956
11 BANDLER, J. W., and CHARALAMBOUS, C: 'Theory of generalized least Pth approxi-
mation', IEEE Trans., CT-19, 1972, pp. 287-289
12 BARBIERE, D.: 'A method for calculating the current distribution of Tschebyscheff
arrays', Proc. IRE, 40, 1952, pp. 78-82
13 BEGOVICH,N. A.: 'Frequency scanning', in Hansen, R. C. (Ed.): 'Microwave scanning
antennas: Vol. III'. Academic Press, 1966, Chap. 2
14 BENNETT, W. R.: 'Methods of solving noise problems', Proc. IRE, 44, 1956, pp. 609-638
15 BICKMORE, R. W., and SPELLMIRE, R. J.: 'A two-parameter family of line sources',
Rept. TM 595, Hughes Aircraft Co., Culver City, California, 1956
16 BLACKMAN, R. B., and TUKEY, J. W.: 'Measurement of power spectra', Dover Publ.,
1958.
17 BLOCH, A., MEDHURST, R. G., and POOL, S. D.: 'A new approach to the design of
super-directive aerial arrays,' Proc. IEE, 100, Part III, 1953, pp. 303-314
18 BLOCH, A., MEDHURST, R. G., and POOL, S. D.: 'Superdirectivity,' Proc. IRE, 48, I960,
p.1164
19 BOUWKAMP, C. J. and DE BRUIJN, N. G.: 'The problem of optimum antenna current
distribution', Philips Research Reports, 1, 1946, p. 135
20 BREITHAUPT, R. W.: 'Conductance data for offset series slots in stripline,' IEEE Trans.,
MTT-16, 1968, pp. 969-970
21 BRICOUT, P. A.: 'Pattern synthesis using weighted functions,' IRE Trans., AP-8, 1960,
pp.441-444
22 BROWN, J. L.: 'A simplified derivation of the Fourier coefficients for Chebyshev pat-
terns', Proc. IEE, 105C, 1957, pp. 167-168
23 BROWN, J. L.: 'On the determination of excitation coefficients for a Tchebycheff pat-
t e r n , ' / / ^ Trans., AP-10, 1962, pp. 215-216
24 BROWN, L. B., and SCHARP, G. A.: 'Tschebyscheff antenna distribution, beamwidth,
and gain tables'. NOLC Report 383, Feb. 1958
25 BUTLER, CM., et al.: 'Characteristics of a wire biconical antenna,' Microwave J., 22,
Sept. 1979, pp. 37-40
26 CARVER, K. R.: 'Practical analytical techniques for the microstrip antenna', Printed Cir-
cuit Antenna Technology Workshop, New Mexico State Univ., 1979
27 CHENG, D.K.: 'Optimization techniques for antenna arrays', Proc. IEEE, 59, 1971, pp.
1664-1674
28 CHESTON, T. C, and FRANK, J.: 'Array antennas', in Skolnik, M. I. (Ed.): 'Radar hand-
book'. McGraw-Hill, 1970
Linear arrays 135

29 CHIGNELL, R. J., and ROBERTS, J.: 'Compact resonant slot for waveguide arrays', Proc.
IEE, 125, 1978, pp. 1213-1216
30 CHU, L. J.: 'Physical limitations of omnidirectional antennas,'/ Appl. Phys., 19, 1948,
pp.1163-1175
31 COURANT, R.: 'Differential and integral calculus: Vol. II.' Interscience Publishers, Inc.,
New York, 1936
32 CRONEY, J., and FOSTER, D.: 'New techniques in the construction of frequency-
scanning arrays',Microwave J. 7, May 1964, pp. 72-74
33 DAS, B. N., and SANYAL, G. S.: 'Network parameters of a waveguide broad wall slot
radiator,' Proc. IEE, 117, 1970, pp. 41-44
34 DERNERYD, A. G., and LIND, A. G.: 'Cavity model of the rectangular microstrip
antenna', Proc. Printed Circuit Antenna Technology Workshop, October 1979, New
Mexico State University
35 DERNERYD, A. G., and LIND, A. G.: 'Extended analysis of rectangular microstrip reso-
nant antennas,' IEEE Trans., AP-27, 1979, pp. 846-849
36 DESCHAMPS, G., and SICHAK, W.: 'Microstrip microwave antenna'. Proc. 1953 AHerton
Antenna Symposium, University of Illinois
37 DIAMOND, B. L.: 'A generalized approach to the analysis of infinite planar array anten-
nas', Proc. IEEE, 56,1968, pp. 1837-1851
38 DIAMOND, B. L., and KNITTEL, G. H.: 'A new procedure for the design of a waveguide
element for a phased-array antenna', in Oliner and Knittel (Eds.): 'Phased array anten-
nas', Artech House, 1972, pp. 149-156
39 DOLPH, C. L.: 'A current distribution for broadside arrays which optimizes the relation-
ship between beam width and side-lobe level,' Proc. IRE, 34, 1946, pp. 335-348
40 DRANE, C.J.: 'Derivation of excitation coefficients for Chebyshev arrays', Proc. IEE,
110,1963, pp. 1755-1758
41 DRANE, C.J.: 'Dolph-Chebyshev excitation coefflceint approximation', IEEE Trans.,
AP-12, 1964, pp. 781-782
42 DRANE, C.J.: 'Useful approximations for the directivity and beamwidth of large scan-
ning Dolph-Chebyshev arrays', Proc. IEEE, 56, 1968, pp. 1779-1787
43 DUHAMEL, R. H.: 'Optimum patterns for endfire arrays', Proc. IRE, 41, 1953, pp. 652-
659
44 ELLIOTT, R. S.: 'Mechanical and electrical tolerances for two-dimensional scanning
antenna arrays', IRE Trans., AP-6, 1958, pp. 114-120
45 ELLIOTT, R. S.: 'Design of line-source antennas for sum patterns with sidelobes of indi-
vidually arbitrary heights', IEEE Trans., AP-24, 1976, pp. 76-83
46 ELLIOTT, R. S.: 'On discretizing continuous aperture distributions', IEEE Trans., AP-25,
1977,pp. 617-621
47 ELLIOTT, R. S.: 'On the design of traveling-wave-fed longitudinal shunt slot arrays',
IEEE Trans., AP-27, 1979, pp. 717-720
48 ELLIOTT, R. S., and JOHNSON, R. M.: 'Experimental results on a linear array designed
for asymmetric sidelobes', IEEE Trans., AP-26, 1978, pp. 351-352
49 ELLIOTT, R. S. and KURTZ, L. A.: 'The design of small slot arrays', IEEE Trans., AP-26,
1978,pp. 214-219
50 EVANS, J.E.: 'Synthesis of equiripple sector antenna patterns', IEEE Trans., AP-24,
1976,pp.347-353
51 EVANS, R. J. and FORTMANN, T. E.: 'Design of optimal line-source antennas', IEEE
Trans., AP-23, 1975, pp. 342-347
52 FLETCHER, R.: 'A new approach to variable metric algorithms', Comp. J., 13,1970, pp.
317-322
53 FLETCHER, R., and POWELL, M. J. D.: 'A rapidly convergent descent method for mini-
mization', Comp. J., 6, 1963, pp. 163-168
136 Linear arrays
54 FLETCHER, R., and REEVES, C. M.: 'Function minimization by conjugate gradients,'
Comp. /., 7, 1964, pp. 149-154
55 FORBES, G. R.: 'An endfire array continuously proximity-coupled to a two-wire line',
IEEE Trans., AP-8, 1960, pp. 518-519
56 FRANZ, K.: 'The gain and the (Rudenberg) 'absorption surfaces' of large directive arrays,'
Hochfrequenztech. u. Elektroakust., 54, 1939, p. 198
57 FRIEDMAN, B.: 'Principles and techniques of applied mathematics'. Wiley, 1956
58 FRIIS, H. T. and LEWIS, W. D.: 'Radar antennas', BSTJ, 26, 1947, pp. 219-317
59 FRY, D. W., and GOWARD, F.K.: 'Aerials for centimetre wavelengths'. Cambridge Uni-
versity Press, 1950
60 GALINDO, V., and WU, C. P.: 'Dielectric loaded and covered rectangular waveguide
phased arrays', BSTJ, 47, 1968, pp. 93-116
61 GANTMACHER, F. R.: 'The theory of matrices: vol. 1.: Chelsea Park, 1960
62 GILBERT, E. N., and MORGAN, S. P.: 'Optimum design of directive antenna arrays sub-
ject to random variations', BSTJ, 34,1955, pp. 637-663
63 GILL, P. E., and MURRAY, W.: 'Quasi-Newton methods for unconstrained optimization',
/. Inst Math. Appl.f 9, 1972, pp. 91-108
64 HADGE, E.: 'Compact top wall hybrid junction', IRE Trans., MTT-1, 1953, pp. 29-30
65 HALL, P. S., et ah'. 'Wide bandwidth microstrip antennas for circuit integration', Elec-
tron. Lett., 15, 1979, pp. 458-460
66 HAMMERSTAD, E. O.: 'Equations for microstrip circuit design'. Proc. European Micro-
wave Conf., 1975, pp. 268-272
67 HANSEN, R. C : 'Gain limitations of large antennas', IRE Trans., AP-8, 1970, pp. 490-
495; see also correction: AP-13, 1965, p. 997
68 HANSEN, R. C: 'Communications satellites using arrays', Proc. IRE, 49, 1961, pp. 1067-
1074
69 HANSEN, R. C : 'Microwave scanning antennas: Vol. 1.' Academic Press, 1964, Chap. 1
70 HANSEN, R. C. (Ed.): Spec. Issue on 'Electronic scanning', Proc. IEEE, 56, 1968, pp.
1761-2038
71 HANSEN, R. C : 'Formulation of echelon dipole mutual impedance for computer', IEEE
Trans., AP-20, 1972, pp. 780-781
72 HANSEN, R. C: 'Significant phased array papers'. Artech House, 1973
73 HANSEN, R. C. and BRUNNER, G.: 'Dipole mutual impedance for design of slot arrays',
Microwave J, 22, 1979, pp. 54-56
74 HANSEN, W. W., and WOODYARD, J. R.: 'A new principle in directional antenna design',
Proc. IRE, 26, 1938, pp. 333-345
75 HARRINGTON, R. F.: 'Effect of antenna size on gain, bandwidth and efficiency', /. Res.
NBS, 64D, 1960, pp. 1-12
76 HARRINGTON, R. F.: 'Sidelobe reduction by nonuniform element spacing', Trans. IRE,
AP-9, 1961, pp. 187-192
77 HARRINGTON, R. F.: 'Time-harmonic electromagnetic fields', McGraw-Hill, 1961
78 HILDEBRAND, F. B.: 'Introduction to numerical analysis', McGraw-Hill, 1956
79 HILL, D. R.: 'Circularly polarized radiation from narrow wall slots in rectangular wave-
guide', Electron. Lett., 16,1980, pp. 559-560
80 HOCKHAM, G. A., and WOLFSON, R. I.: 'Frequency scanning antenna using evanescent-
mode waveguide', Proc. IEE Conf. on Antennas, Nov. 1978, pp. 21-24
81 HYNEMAN, R. F.: 'Closely-spaced transverse slots in rectangular waveguide',//?£* Trans.,
AP-7, 1959, pp. 335-342
82 INCE, W. J., and TEMME, D. H.: 'Phasers and time delay elements', in Young, L., 'Advan-
ces in microwaves: Vol. 4'. Academic Press, 1969, pp. 1-189
83 ISHIMARU, A.: 'Theory of unequally-spaced arrays', IEEE Trans., AP-10, 1962, pp. 691-
702
Linear arrays 137

84 JAECKLE, W. G.: 'Antenna synthesis by weighted Fourier coefficients', IEEE Trans.,


AP-12, 1964, pp. 369-370
85 JAMES, J. R., and HALL, P. S.: 'Microstrip antennas and arrays. Pt. 2: New array-design
technique', IEE J. Microwaves, Optics & Acoustics, 1, 1977. pp. 175-181
86 JAMES, J. R., and HENDERSON, A.: 'High-frequency behavior of microstrip open-circuit
terminations', ibid. 3, 1979, pp. 205-218
87 JORDAN, E. C, and BALMAIN, K. G.: 'Electromagnetic waves and radiating systems'.
Prentice-Hall, Inc., New Jersey, 1968, Section 14.08
88 KAMINOW, I. P., and STEGEN, R. J.: 'Waveguide slot array design'. TM-348, Hughes Air-
craft Co., Culver City, Cal., July 1954
89 KERR, J. L.: 'Microstrip antenna developments', Proc. Printed Circuit Antenna Tech-
nology Workshop, October 1969, New Mexico State University
90 KHAC, T. V., and CARSON, C. T.: 'Impedance properties of a longitudinal slot antenna
in the broad face of a rectangular waveguide*, IEEE Trans., AP-21, 1973, pp. 708-
710
91 KHAC, T. V.: 'A study of some slot discontinuities in rectangular waveguides'. Ph.D.
thesis, Monash University, 1974
92 KING, D. D., PACKARD, R. F., and THOMAS, R. K.: 'Unequally-spaced, broad-band
antenna arrays',/JEW-F Trans., APS, I960, pp. 380-384
93 KING, H. E. and WONG, J. L.: 'An experimental study of a balun-fed open-sleeve di-
pole in front of a metallic reflector', IEEE Trans., AP-20, 1972, pp. 201-204
94 KING, R. W. P.: 'The theory of linear antennas'. Harvard University Press, 1956
95 KNITTEL, G. H.: 'Wide-angle impedance matching of phased-array antennas - A Survey
of Theory and Practice' in Oliner and Knittel (Eds.): 'Phased Array Antennas'. Artech
House, 1972, pp. 157-172
96 KNITTEL, G. H., et al.: 'Element pattern nulls in phased arrays and their relation to
guided waves', Proc. IEEEf 56, 1968, pp. 1822-1836
97 KSIENSKI, A.: 'Maximally flat and quasi-smooth sector beams', IRE Trans., AP-8, 1960,
pp. 476-484
98 KUMMER, W. H.: 'Feeding and phase scanning' in Hansen, R. C. (Ed.): 'Microwave scan-
ning antennas'. Academic Press, 1966, Vol. Ill
99 KURTZ, L. A., and ELLIOTT, R. S.: 'Systematic errors caused by the scanning of antenna
arrays: Phase shifters in the branch lines', IRE Trans., AP-4,1956, pp. 619-627
100 LAPAZ, L., and MILLER, G. A.: 'Optimum current distributions on vertical antennas',
Proc. IRE, 31, 1943, pp. 214-232
101 LEE, S. W. and MITTRA, R.: 'Radiation from dielectric-loaded arrays of parallel-plate
waveguides', IEEE Trans., AP-16, 1968, pp. 513-519
102 LEVY, R.: 'Directional couplers', in Young, L. (Ed.): 'Advances in microwaves: Vol. 1.'
Academic Press, 1966, pp. 115-209
103 LO, Y. T.: 'A note on the cylindrical antenna of noncircular cross-section', /. AppL Phys.,
24,1953,pp.1338-1339
104 LO, Y. T.: 'A mathematical theory of antenna arrays with randomly spaced elements',
IEEE Trans., AP-12, 1964, pp. 257-268
105 LO, Y. T., LEE, S. W. and LEE, Q. H.: 'Optimization of directivity and signal-to-noise
ratio of an arbitrary antenna array,' Proc. IEEE, 54, 1966, pp. 1033-1045
106 LO, Y. T. et aL: 'Theory and experiment on microstrip antennas', IEEE Trans., AP-27,
1979,pp. 137-145
107 MAGILL, E. G. and WHEELER, H. A.: 'Wide angle impedance matching of a planar array
antenna by a dielectric sheet', IEEE Trans., AP-14, 1966, pp. 49-53
108 MARCUM, J. I.: 'Studies of target detection by a pulsed radar-mathematical appendix',
IEEE Trans., IT-6, 1960, pp. 159-160 and 227-228
109 MAUTZ, J. R. and HARRINGTON, R. F.: 'Radiation and scattering from loaded bodies
of revolution', IEEE Trans., AP-23, 1975, p. 594
138 Linear arrays
110 McGEE, W. F.: 'Another recursive method of computing the Q function', IEEE Tram.,
IT46,1970, pp. 500-501
111 METZLER, T.: 'Microstrip series arrays', Proc. Printed Circuit antenna Technology Work-
shop, October 1979, New Mexico State University
112 MUCCI, R. A. et al: 'Beam pattern synthesis for line arrays subject to upper and lower
constraining bounds', IEEE Trans., AP-23,1975, pp. 732-734
113 NEWMAN, E. H. et al: 'Superdirective receiving arrays', IEEE Trans., AP-26, 1978, pp.
629-635
114 OLINER, A. A.: 'The impedance properties of narrow radiating slots in the broad face of
rectangular waveguide', IRE Trans., AP-5, 1957, pp. 4-20
115 OLINER, A. A. and MALECH, R. G.: 'Radiating elements and mutual coupling', in
Hansen, R. C. (Ed.): 'Microwave Scanning Antennas: Vol. 2.: Academic Press, 1966
116 OLINER, A. A., and MALECH, R. G.: 'Mutual coupling in infinite scanning arrays', in
Hansen, R. C. (Ed.): Microwave scanning antennas: Vol. 2.' Academic Press, 1966
117 OSEEN, C. W.: 'Die Einsteinsche Nadelstichstrahlung und die Maxwellschen Gleichungen,'
Ann. derPhysik, 69, 1922, p. 202
118 PARK, P. K., and ELLIOTT, R. S.: 'Design of colinear longitudinal slot arrays fed by
boxed stripline', IEEE Trans., AP, to be published
119 PERINI, J., and IDSELIS, M. H.: 'Radiation pattern synthesis for broadcast antennas',
IEEE Trans., BC-18, 1972, pp. 53-62
120 PRITCHARD, R. L.: 'Optimum directivity patterns for linear point arrays,' /. Acoust.
Soc. America, 25, 2953, pp. 879-891
121 REDLICH, R. W.: 'Iterative least-squares synthesis of nonuniformly spaced linear arrays',
IEEE Trans., AP-21, 1973, pp. 106-108
122 REICH, H. B., et al: 'Very high frequency techniques: Vol. L' McGraw-Hill, 1947, Chap.
3
123 REID, D. G.: The gain of an idealized Yagi array,' /. IEE, 93, Part IIIA, 1946, pp. 564-
566
124 RHODES, D. R.: The optimum line source fo the best mean-square approximation to a
given radiation pattern,' IEEE Trans. AP-11, 1963, pp. 440-446
125 RHODES, D. R.: 'On an optimum line source for maximum directivity', IEEE Trans.
AP-71, 1971, pp. 485-492
126 RHODES, D. R.: 'Synthesis of planar antenna sources', Clarendon Press, 1974
127 RIBLET, H. J.: 'Discussion on 'A current distribution for broadside arrays which opti-
mizes the relationship between beam width and side-lobe level', Proc. IRE, 35, 1947,
pp.489-492
128 RIBLET, H. J.: 'Short-slot hybrid junction', Proc. IRE, 40, 1952, pp. 180-184
129 RONDINELLI, L. A.: 'Effects of random errors on the performance of antenna arrays of
many elements'. IRE Nat. Conv. Rec, Pt. 1, 1959, pp. 174-189
130 ROTHMAN, M.: Table of flo(x)dx for 0(0.1)20(1)25'. Quart. J. Mech. Appl Math., 2,
1949, pp. 212-217
131 RUZE, J.: The effect of aperture errors on the antenna radiation pattern', Nuovo
Cimento, 9 Suppl., 1952, pp. 364-380
132 RUZE, J.: 'Antenna tolerance theory - A review', Proc. IEEE, 54, 1966, pp. 633-640
133 SALZER, H. E.: 'Calculating Fourier coefficeints for Chebyshev patterns', Proc. IEEE,
63,1975,pp. 195-197
134 SANDLER, S. S.: 'Some equivalence between equally and unequally spaced arrays', IEEE
Trans., AP-8, 1960, pp. 496-500
135 SCHELKUNOFF, S. A.: 'A mathemetical theory of linear arrays,' BSTJ, 22, 1943, pp.
80-107
136 SCHIFFMAN, B. M.: 'A new class of broadband microwave 90° phase shifters', IRE
Trans., MTT-6, 1958, pp. 232-237
Linear arrays 139

137 SCHWARZ, S. E., and RUTLEDGE, D. B.: 'Moving toward NMM wave integrated circuits',
Microwave /., 23, June 1980, pp. 47-52
138 SCHIMZU, J. K., and JONES, £. M. T.: 'Coupled-transmission line directional couplers',
IRE Trans., MTT-6, 1958, pp. 403-410
139 SKOLNOK, M. L: 'Introduction to radar systems'. McGraw-Hill, 1962
140 SLATER, J. C: 'Microwave electronics'. Van Nostrand, 1950
141 SOUTHWORTH, G. C: 'Arrays of linear elements', in Jasik, H. (Ed.): 'Antenna engineer-
ing handbook'. McGraw-Hill, 1961
142 SPELLMIRE, R. J.: 'Tables of Taylor aperture distributions', Rept. TM581, Hughes Air-
craft Co., Culver City, Cal., 1958
143 SPRADLEY, J. L., and ODLUM, W.J.: 'Systematic errors caused by the scanning of
antenna arrays: Phase shifters in the main feed line'. Rept. AFCRL-56-795, SR11/
1317, Hughes Aircraft Co., Culver City, Cal., 1956
144 STARK, L. et al.\ 'Multi mode phased array element for scan angle compensation of ele-
ment impedance', Alta Freq., XXXVIII, 1969, pp. 249-254
145 STARK, P. A.: 'Introduction to numerical methods', Macmillan PubL, 1970
146 STEGEN, R. J.: 'Excitation coefficients and beamwidths of Tschebyscheff arrays,' Proc.
IRE, 41, 1953, pp. 1671-1674
147 STEINBERG, B. D.: 'Comparison between the peak sidelobe of the random array and
algorithmically designed aperiodic arrays', IEEE Trans., A-21, 1973, pp. 366-370
148 STEINBERG, B. D.: 'Principles of aperture and array system design', Wiley, 1976
149 STEVENSON, A. F.: 'Theory of slots in rectangular waveguides',/. AppL Phys., 19,1948,
pp.24-38
150 STUTZMAN, W. L.: 'Synthesis of shaped-beam radiation patterns using the iterative
sampling method', IEEE Trans., AP-19, 1971, pp. 36-41
151 STUTZMAN, W. L.: 'Sidelobe control of antenna patterns', IEEE Trans,, AP-20,1972, pp.
102-104.
152 STUTZMAN, W. L.: 'Shaped-beam synthesis of nonuniformly spaced linear quays', IEEE
Trans., AP-20, 1972, pp. 499-501
153 TAI, C. T.: 'Characteristics of linear antenna elements' in Jasik, H. (Ed.): 'Antenna engin-
eering handbook', McGraw-Hill, 1961
154 TAI, C.T.: 'The optimum directivity of uniformly spaced broadside arrays of dipoles',
IEEE Trans., AP-12, 1964, pp. 447-454
155 TAMIR, T.: 'Leaky-wave antennas' in Collin, R. E., and Zucker, F. J. (Eds.): 'Antenna
theory: Part 2'. McGraw-Hill, 1969
156 TANG, R., and WONG, N. S.: 'Multimode phased array element for wide scan angle im-
pedance matching', Proc. IEEE, 56, 1968, pp. 1951-1959
157 TAYLOR, T. T.: 'One-parameter family of line sources producing modified su\itu/iru pat-
terns'. Rept. TM 324, Hughes Aircraft Co., Culver City, Cal., 1953
158 TAYLOR, T. T.: 'Design of line-source antennas for narrow beam width and low sidelobes',
IRE Trans., AP-3, 1955, pp. 16-28
159 TEMES, G. S., and ZAI, D. Y. F.: 'Least Pth approximation', IEEE Trans., CT-16, 1969,
pp.235-237
160 TSANDOULAS, G. N.: 'Wideband limitations of waveguide arrays', Microwave J., 15,
Sept. 1972, pp. 49-56
161 TSANDOULAS, G. N., and KNITTEL, G. H.: 'The analysis and design of dual-polarization
square-waveguide phased arrays', IEEE Trans., AP-21, 1973, pp. 796-808
162 TSANDOULAS, G. N.: 'Unidimensionally scanned phased arrays', IEEE Trans., AP-28,
1980, pp. 86-99
163 TSENG, F. I.: 'Design of array and line-source antennas for Taylor patterns with a null',
IEEE Trans., AP-27, 1979, pp. 474-479
164 TSENG, F. I., and CHENG, D. K.: 'Antenna pattern response to arbitrary time signals',
Can. J. Phys., 42, 1964, pp. 1358-1368
140 Linear arrays
165 TYRELL, W. A.: 'Hybrid circuits for microwaves', Proc. IRE, 35, 1947, pp. 1294-1306
166 UNZ, H.: 'Nonuniform arrays with spacings larger than one wavelength', IEEE Trans.,
AP-10, 1962, pp. 647-648
167 UZKOV, A. L: 'An approach to the problem of optimum directive antennae design', C.R.
Acad. Sci. URSS., 53, 1946, p. 35
168 UZSOKY, M., and SOLYMAR, L.: 'Theory of super-directive linear arays,' Acta Phys.
Hung., 6, 1956, pp. 185-205
169 VAN DER MAAS, G. J.: 'A simplified calculation for Dolph-Tchebycheff arrays', JAP,
25,1954,pp.121-124
170 VOGES, R. C, and BUTLER, J. K.: 'Phase optimization of antenna array gain with con-
strained amplitude excitation', IEEE Trans., AP-20, 1972, pp. 432-436
171 WASYLKIWSKYJ, W., and KAHN, W. J.: 'Element pattern bounds in uniform phased
arrays*, IEEE Trans., AP-25, 1977, pp. 597-604
172 WATERMAN, A.: 'Stripline antenna development', Proc. Printed Circuit Antenna Tech-
nology Workshop, October 1979, New Mexico State University
173 WATSON, W. H.: 'The physical principles of waveguide transmission and antenna systems',
Oxford University Press, 1947
174 WESTLAKE, J. R.: 'A handbook of numerical matrix inversion and solution of linear
equations'. Wiley, New York, 1968, p. 32
175 WHEELER, H. A.: 'Simple relations derived from a phased array antenna made of an
infinite current sheet', IEEE Trans., Ap-13, 1965, pp. 506-514
176 WHITTAKER, E. T., and WATSON, G. N.: 'Modern analysis'. Cambridge University Press,
1952,4th Edn.
177 WILLEY, R. E.: 'Space tapering of linear and planar arrays', IEEE Trans., AP-10, 1962,
pp. 369-376
178 WINKLER, L. P. and SCHWARTZ, M.: 'A fast numerical method for determining the
optimum SNR of an array subject to a Q factor constraint*, IEEE Trans., AP-20,
1972, pp. 503-505
179 WONG, N. S., et ai: 'Multimode phased array element for wide scan angle impedance
matching', in Oliner and Knittel (Eds.): 'Phased Array Antennas', Artech House, 1972,
pp. 178-186
180 WOODWARD, P. M.: 'A method of calculating the field over a plane aperture required to
produce a given polar diagram', Proc. IEE., 93 Pt. Ill, 1947, pp. 1554-1558
181 WOODWARD, P. M. and LAWSON, J. D.: 'The theoretical precision with which an arbit-
rary radiation pattern may be obtained from a source of finite size', Proc. IEE, 95 Pt.
Ill, 1948, pp. 363-370
182 WU, C. P. and GALINDO, V.: 'Surface-wave effects on dielectric sheathed phased arrays
of rectangular waveguides', BSTJ, 47, 1968, pp. 117-142
183 YARU, N,: 'A note on supergain antenna arrays,' Proc. IRE, 39, 1951, pp. 1081-1085
184 YEE, H. Y.: 'Impedance of a narrow longitudinal shunt slot in a slotted waveguide array',
IEEE Trans., AP-22, 1974, pp. 589-592
185 YOUNG, L. B.: in Montgomery, C. G. (Ed.): 'Technique of microwave measurements'.
MIT Radiation Lab. Series, Vol. 11, McGraw-Hill, 1947
186 YOUNG, L.: 'Parallel coupled lines and directional couplers'. Artech House, 1972
Chapter 10

Planar arrays
R. C. Hansen

Introduction

Planar arrays and linear arrays have much in common; thus many aspects of separ-
able arrays, i.e. those rectangular arrays with separable x and y excitations, have
been covered in Chapter 9. Included here are features that are specific to a planar
array such as change of beam shape with scan angles, grating and quantisation lobes,
excitations, directivity and, most important, feeding methods. Mutual coupling is
the key to understanding array behaviour as a function of element spacing, scan
angles etc., so it is the subject of a major part of the Chapter. Linear-array mutual
coupling effects are included here as they were omitted in Chapter 9.
Section 10.1 discusses array beam shape and beam width vs. scan, rectangular
and hexagonal lattices and their grating and quantisation lobes. The latter arise
from use of digital phasers. Directivity formulations are followed by discussion of
two circular (non-separable) aperture distributions that are used directly for large
arrays, and as a starting function for small arrays: the Taylor h and the Hansen one-
parameter. Feeds are the subject of Section 10.2, with fixed beam, single plane of
scan, and full-scan arrangements discussed. Finally mutual coupling, which is both a
difficulty to be reckoned with, and a means for understanding the subtleties of
arrays, occupies Section 10.3. An attempt is made to relate all recent concepts and
theories, including active impedance and active element pattern, isolated and
minimum scattering elements, the grating lobe series, active element efficiency, the
periodic cell, blind angles, matching networks in frequency, and matching networks
in spatial frequency. A unified understanding is now emerging.
Again it is a pleasure to thank many colleagues for help and counsel.

10.1 Array excitation

10.1,1 Array lattice


Spherical coordinates with the polar axis normal to the array plane are convenient
for planar arrays (see Fig. 10.1). The array plane \%x-y. Regular arrays of elements
142 Planar arrays

usually utilise a rectangular or triangular lattice; Fig. 10.1 shows the former. For
example, a rectangular array of waveguide slots may use half-guide wavelength
along one axis and outside-guide-width spacing along the other axis. Sometimes
alternate rows of slots are staggered to form an isosceles triangular lattice. A circular
disc array may use a rectangular lattice or a staggered triangular lattice. Sometimes
an equilateral triangular (hexagonal) lattice is used (see Fig. 10.2). Hexagonal lattices
are discussed below under grating lobes.

Fig. 10.1 Spherical coordinate system and rectangular lattice

O C)

o o o

e e e

o o

Oy JO O 0 O O O

V
elements

Fig. 10.2 Triangular lattice


Asymmetric excitations are not used except for shaped beams, so for simplicity
only symmetric excitations are considered in this chapter. An even number of
elements N along x, andM along y, is assumed. The space factor of a planar array is
then:
Planar arrays 143

N/2M/2 {2n-\)kdxu (2m-\)kdyLv


F(u,v) — 2 - 2 - >l Rm cos cos ' — (10.1)

Here the direction cosine plane variables are:


u = sin 6 cos 0 — sin 60 cos 0O
v = sin d sin 0 — sin d0 sin 0O (10.2)
e
with #o, 0o th beam pointing angles. Element lattice spacings are dX9 dy and k —
2TT/X. Interelement phase shift needed to position the beam (assuming a pencil
pattern) is:
$u - kdxu0 = kdx sin 60 cos 0O
<£>v = kdyVo =fo/ysin #o sin 0O (10.3)
In a later Section various ways of producing these phase shifts in a planar array will
be discussed.

10.1.2 Beamwidth
To calculate beamwidth of an array exactly, a numerical procedure must be used.
Since most tapered excitations derive array coefficients by sampling a continuous
distribution, it is appropriate to calculate beamwidth from continuous distributions.
Take a rectangular array with separable distribution along x and y, each of which is
a Taylor one-parameter (see Chapter 9). Then the space factor is:
F(u,v) = sine TI\JU2 — B2 sine n\Jv2 — B2 (10.4)
Here the Taylor w, v are used, and L and W are array length and width:
L ,
u = — (sin$ cos0— sin 6o cos0 o )
X
w
v - — (sin 6 cos 0 — sin 60 cos 0O) (10.5)
X
Only scan in the x plane (0 = 0) will be considered, as principal-plane scan is of
most interest. Taking first the beamwidth in the scan plane where v = 0:
sinh nB r-~ = L, ^ , ^^
F(u, v) = — sine iry/u2~B2y u = - ( s i n 0 - sin 0O) (10.6)
TtJj A

Since the value on axis is (sinh KB/KB)2, the beamwidth equation is:
sinh nB
(10.7)

In this form it is assumed that u3 > B, which occurs when

. „_ or B< 0.4747 (10.8)


nB
144 Planar arrays

This corresponds to a sidelobe ratio of 16.27 dB. Thus for lower SLR (higher side-
lobes) including uniform excitation, eqn. 10.7 is correct. For higher SLR the proper
equation is:

(10.9)
\phiB
These equations can be solved numerically given B, and the results are exactly those
given in Section 9.1.1 for linear arrays. Of course when u3 is determined, the actual
3 dB beamwidths are found from the scan angle 0O using:

63x = arc sin I — - sin 0O) 4-arc sin I — + sin 0O) (10.10)

As the beam moves towards endfire, the inside 3 dB point will occur at 90°; this
happens when sin0 o + u3X/L = 1. Beyond this the definition of beamwidth must
be changed, and to be consistent at endfire it should be the angle between the outer
3 dB points of the conical pattern. Thus at the transition angle, the beamwidth
abruptly doubles. For sin 0O > 1 — u3\jL:

63x = 7T+ 2 arc sin ( — - s i n 0ol (10.11)

Fig. 10.3 shows scan-plane beamwidth for square apertures of various lengths.
Cross-plane beamwidth is more difficult as both L and W affect it. Let 0O = 0
and 0 = 0O. Then:
L
u = — sin 0 o (cos 0O— 1)
A
W
v = —sin 0O sin 0 (10.12)
A
For large sidelobe ratios the half power u3,v3 are solutions of:
sinh TI\JB2 — U\ sinh iwjB1 — v\ sinh2KB
= (10.13)
" V2B2
Given L, W and B, a value of 0 3 can be found that satisfies the equation. If u3 or
v3 is greater than B, the hyperbolic form changes to trigonometric. The cross-plane
beamwidth is found from:
0 3 y = 2 arc sin (sin 0O sin 0 3 ) (10.14)
where the sin 0O factor represents a projected aperture. When the array is large, an
excellent approximation is:
0 3 y ~ 2 sin 0O sin 0 3 (10.15)
It is convenient to use the array aspect ratio W/L. For a given W/L andi?, the root
Planar arrays 145
0 3 can be determined as a function of (W/X) sin 60. Then the cross-plane beam width
can also be written in terms of v3:
W
— (10.16)
A

100 r

10

0-1
10 20 30 40 50 60 70 80 90
scan angle, deg

Fig. 10.3 Square uniform aperture. Beam width in scan plane

So cross-plane beamwidth, normalised by W/\y can be plotted vs. in-plane scan


angle, also normalised by W/X. Figs. 10.4 and 10.5 give these results for uniform
excitation and for 25 dB SLR.
146 Planar arrays

60 r

50

g 30
0 5

20

10

0-2 0-4 06 1 2 U 6 10

Fig. 10.4 Uniform rectangular aperture. Beamwidth per wavelength in plane normal to scan

70

60

50

"9 U0
E

' 30- 0-5

20-

0-2 04 06

Fig. 10.5 25dB SLR rectangular aperture. Beamwidth per wavelength in plane normal to scan
Planar arrays 147

As the beam is scanned its shape changes as the projected aperture of the array
changes. For rectangular arrays the 3 dB beam contour is approximately elliptical.
However, for scans not in the principal planes, the combination of projected aperture
width and length and the scan angles results in an elliptical beam whose major dia-
meter is generally not oriented along the scan plane or principal planes (Elliott 25 ).
Since the area of an ellipse is proportional to the product of major and minor dia-
meters, the 3 dB beam area, called areal beamwidtb, is proportional to the product
of the major- and minor-axis beamwidths. The areal beamwidth is to first order
independent of azimuth angle 0, although the beam shape and orientation of major
axis may change with 0. Fig. 10.6 shows several beams as they change with scan.

Fig. 10.6 Pencil beam shape with scan {courtesy: Elliott)

10.1.3 Grating lobes

Rectangular lattice: The array lattice controls the appearance of grating lobes.
Each point on the lattice contains an array element. A rectangular lattice with scan-
ning in either principal plane behaves exactly like a linear array (see Chapter 9), in
grating-lobe behaviour. At other scan angles the situation is not so simple, but can
be easily analysed by plotting the positions of the GL in the u, v plane* as developed
by Von Aulock.94 As shown in Fig. 10.7 the main beam and grating lobes occur at
the points of the inverse lattice; that is, the lattice spacing is \/dx and X/dy. All real
angles, sometimes called visible space, are inside or on the unit circle, and, of
course, when 6 = ir/2, u2 + v2 — 1. Angles outside the unit circle are then imaginary,
or in invisible space. When the main beam is scanned, the origin moves to the new
value of u, v, and all grating lobes move correspondingly. Thus, for scan in the
w-plane (0 = 0) the main beam point moves toward 4- 1 for 6 > 0 and all GL points
move the same amount parallel to the w-axis. When the GL just left of the unit
circle moves enough to intersect the unit circle, that GL becomes visible. Thus the

* This is sometimes called the direction cosine plane


148 Planar arrays

distance between the broadside GL and the unit circle must be no larger than unity
or the GL will not intersect the unit circle before the main beam stops at the right
side of the unit circle. The result is eqn. 9.10 of Chapter 9:

X dx
= 0 (10.17)
1 or
X sin
Unlike Chapter 9 these results have the inequality arranged to allow the grating
lobe to appear. Clearly scan along v(<j> = ir/2) is analogous:

0 = 7T/2 (10.18)
X "" 1 + sin B'

or cr T. S° </> a^

Fig. 10.7 Grating lobes in u, v plane

Of course scan to any pair of angles, or pair of u, v9 will produce one or more
GL if the array spacings are sufficiently large. The limiting cases are of more
interest, and these are of two types. First is diagonal plane scan sketched in Fig.
10.8. This is the GL produced when the diagonal point intersects the unit circle
normally. Values of u, v that allow this GL to appear are:
SQ-l SQ-l
(10.19)
SQdyl\

$in6gl = SQ-l, tan0,, = -f (10.20)


Cly

where SQ = \J\2/d% + X2/d^. This diagonal plane GL only occurs for SQ>2. In
Planar arrays 149

general, a grating lobe appears whenever the propagation constant is real; at the
transition from imaginary to real, it is zero.

- = \ / l - (u - nXjdxf + (v - mk/dy)2 = 0 (10.21)


K

tangent
GL

Fig. 10.8 Grating lobe formation off axes

The diagonal case corresponds to n = 1, m = 1. When spacings are sufficiently large


to allow several GL to appear, the propagation constant is real for larger n and/or
m. Of course, scan in principal planes corresponds to n = 0 or m — 0. The other
limiting case is when the GL located on the w-axis just left of the unit circle scans to
be just tangent to the circle (see Fig. 10.8). Clearly only dx is involved here. The
limiting values are:

vgl = V i - (10.22)
dj\

sin< sin0 = dxjX (10.23)


dJX
The minimum value of dx/X for this lobe is l/%/2. The v tangent case is analogous
as might be expected, with w, v and dx, dy interchanged. For unrestricted scan the
principal plane GL appears first, controlled by the smallest ofX/dx and X/dy. Thus
the u, v plane not only gives an excellent physical picture of grating-lobe occur-
rence, but also allows the formulas to be easily derived.
It should be emphasised that if dJX < 1/(1 + sin d) and dy/X < 1/(1 + sin 0),
no grating lobes will exist for the rectangular lattice.

Hexagonal lattice: A rectangular lattice is not necessary; an isosceles triangular


lattice can be used instead. A special case that is commonly used is the equilateral
triangular lattice, sometimes called the regular hexagonal lattice. If the array
150 Planar arrays

elements are all equidistance by 2d, and the x-axis goes through a row of elements
spaced this distance apart, dx = d and dy = \[%d\2 where dx and dy are half the
distance to the next element, along the x or y axis. The hexagonal array is easily
analysed by separating it into two interlaced rectangular lattice arrays. The pattern
is then the sum of the two array patterns, and the grating-lobe locations for no
beam scan are simply:
nk mX mX
ugl = + • (10.24)
'Id'
The advantages of the hexagonal lattice will appear later. Fig. 10.9 shows the
inverse lattice in the w, v plane, where the GL points are all equidistant by \jd.

Fig. 10.9 Grating-lobe limiting cases

Scan in the w-plane is as before. Diagonal plane scan normal to the unit circle gives a
GL for:
\/3(\-d/X)
u8l = (10.25)
2d/X 2d/X
l-d/X
sin Uai = <t>gi = 6 0 ° . (10.26)
d/X
The minimum value of d/X for this lobe to appear is \. There are two tangent cases,
where the GL normally on the w-axis becomes tangent to the unit circle, and where
the normally diagonal GL becomes tangent. The first occurs for:
Planar arrays 151

l-d2jX2
ugl = ^ , vgl = V l ~d2i\2 (10.27)

sin 0,| = : , sin0,, = d/X (10.28)


d/X
The minimum value of d/X for this GL to appear is l/\/2.
Tangency from the diagonal point is somewhat more complicated and occurs
for:
y/l-d2l\*{y/\-dV\2

Vi-rf2/x2A/3(i-<W)
• i

yj\-d2jtf n d
= — , <pgl = - - a r c s i n - (10.30)
uj k 3 A

In comparing the hexagonal lattice with the square lattice, both have grating-lobe
appearance when sin 6 = X/d — 1. However the element area in a square lattice is
d2, while that in the hexagonal lattice is 2d2\\Jl>. Thus the hexagonal lattice
requires only 0.866 as many elements to give the same grating-lobe-free area, or a
saving of 15% in number of elements (Sharp;86 Lo and Lee64).

10.1.4 Quan tisation lobes


In Chapter 9 there is an extensive discussion of quantisation lobes (QL) produced
by digital phasers. For rectangular lattices the results derived there are valid in
general, except that the QL tend to appear in bands in w, v space (see Fig. 10.10).
In this Figure a square lattice is shown. The bands indicate where higher sidelobes
due to phaser quantisation will appear, and of course only those bands within
the unit circle represent loci of visible QL. When a hexagonal lattice is used with a
hexagonal phaser network (a two-dimensional structure) the QL bands follow the
principal diagonals, and in most cases only two bands would cross the unit circle.
Hexagonal networks require many circuit-board crossovers, and hence are seldom
used. A more common scheme uses row and column phasers, which produce QL
bands parallel to the u and v axes [Nelson73]. Fig. 10.11 shows such a situation
in the w, v plane, and it can be seen that an additional band of quantisation lobes
occur. Thus the hexagonal lattice offers grating lobe advantages, but no quantis-
ation lobe advantages when u-v (row-column) phasing is used.

10.1.5 Directivity
For simplicity a uniformly spaced array disposed on a rectangular lattice, with
broadside beam, is considered. The directivity is given by:
152 Planar arrays

D = (10.31)
F2 ineded(
Jo "Jo * t>
Now insert eqn. 10.1 with u0 = vQ = 0 into this expression. The numerator
immediately becomes:
/iV/2 Mil \2
NUM (10.32)

Fig. 10.10 Quantisation-lobe strips for square lattice

m m m

Fig. 10.11 Quantisation lobe strips for hexagonal lattice


Planar arrays 153
2
In the denominator F becomes a fourfold sum:

Z ZZZ
Z Z AnmApq cos (2n - 1) a cos (2m-1)$ cos (2p-l) a
n m p q
xcos(2<?- 1)0 (10.33)
with a = (ndj\)u and 0 = (irdy/\)v.
Through liberal and frequent use of trigonometric identities the four-cosine pro-
duct becomes a sum of eight terms where the integer factors A and B occur in all
possible combinations. The general term is:
i cos (Aa ± £0)
and the values A and B are:

m +q—

These can be recast into the form:


cos (a cos 0 4- & sin 0) = cos [v# 2 + Z?2 sin (0 4- arc tan a/b)]

This last form can be integrated in 0:

j2j cos [Vtf2 + b2 sin (0 4- arc tan a/b)] dcp = 2nJ0(\/a2 4 6 2 ) (10.34)

Fortunately this result allows the 0 integration to be performed:

2TT JJ/oCVfl2 + ^ 2 ) sin ade = 4TT sine C (10.35)

where \Ja2 + b2 = C sin ^. Finally the directivity can be written:

/JV/2M/2 \2
4 Z I Anm
D = Mil Nil MJ2 \ (10.36)
I
where
?! = sine 2n\J(n 4 p - l)2d2/X2 + (m + q-

= sine 2ny/(n 4 p - 4 (m - q)2d2/\2

53 = sine 2ity/(n-p)2d2/X2 + (m + q - l)2d2lX2

54 = sine 2ny/(n -p)2d2/X2 4 (m-q)2d2/\2 (10.37)


154 Planar arrays

The function sine x = sinx/x. Unfortunately this result for directivity does not
simplify appreciably either for uniform excitation or for half-wave spacing in the
lattice. For the half-wave case S2 simplifies to:

S2 = sine ny/(n + p - I) 2 + (m - qf (10.38)

Only for special sets of integers is S2 = 0 (for example, n 4- p = 4, m — q = 4, so


V = 5). The cross-coupling terms then are not in general zero.
Although the directivity expression is for a rectangular array with rectangular
lattice it can be readily modified for other types. For example, a flat plane array,
where the waveguide linear arrays are of different lengths, is handled by adjusting
the summation limits. Other lattices such as hexagonal can also be accommodated.
In general, it is necessary to use approximations to obtain planar array direc-
tivity. For example, the directivity of a planar array is approximately:

D~2DxDy
where Dx and Dy are the linear array directivities along x and along y. The factor
of 2 gives a result that compares well with exact square-array calculations
(Hansen44); the factor of 2 accounts for a pencil beam radiated on one side only.
In the previous derivation the element was assumed to be isotropic. Directivity
of an array of actual elements can be computed whenever the mutual impedance
between only two of the elements can be calculated. Following the method of
Section 9.1.1.2, the general directivity formula is:

120 I
n m
D = (10.1.39)
2- Ld Lu 2- AnmApqKnmpq
n m p q

Here the sums go over all the nm elements, and Rnmpq is the mutual resistance
between the nmth element and the pqth element. For slots and dipoles fast and
efficient algorithms exist for calculation of mutual resistance (Hansen and
Brunner47) so that calculation of directivity without approximations is facilitated.
A useful picture of the variation of planar array directivity with scan angle can
be obtained by considering a uniformly excited planar aperture. The usual spherical
coordinate system is used but the aperture is in the y-z plane. Thus the aperture
normal is d0 = n/2, <p0 = 0. The space factor is now:

E = sine TTU sine irv (10.40)

where

u = — (cos B — cos # 0 ), v = ~ sin 6 sin 0


A A
Planar arrays 155

Scan in the <j>o = 0 plane is assumed. Directivity is now:

G = — (10.41)

J_ W / J O * ' * » « « » < *
>7T/2 [IT

The 0 integral can be simplified by changing the variable to v. The necessary cos <j>
is inserted since it is essentially unity over the beam (for large apertures), with
larger error at angles where the sidelobe energy is low. Thus the <j> integral is
approximately:
cnl2 n X
E2d(j>~ (10.42)

Directivity is now approximately:

A similar change of variable produces the result (King and Thomas57):

G^4nLW^0o (1044)
X
This result, which is valid as long as the beam is narrow in both planes, states that
the effective aperture varies as sin 0Q. Near endfire the same (j> integral approxi-
mation is used, but the 6 integration must be approximated differently. There,

which gives near endflre:


47r X2 r 1 sin27Tw &u
J 7rW(2Xw/Z)lI
G "Z^Jo (1
'4
At endfire the result is

G = 37rx/2f f \ (10.46)
XV X
Fig. 10.12 shows how the effective aperture varies with principal plane scan, and
from this the range of usefulness of the sin 0o effective aperture may be observed.
The sin 60 effective aperture decays to the endfire value at:

sin 0O = i (10.47)
2y/2L/X
so it should be used only for scan angles well above this value.
156 Planar arrays

-2 -

—__L = 5

"- 10
\

- 6 --
\

COS8 0
-
\

-10 1 1 I 1 1
20 60 80 90
9 0 ,deg

Fig. 10.12 Effective aperture vs. scan angle

10.1.6 Planar array excitation

Taylor circular h distribution: Rectangular arrays can be designed by using linear


distributions along x and y axis; see Chapter 9. For arrays whose boundary is a
circle, commonly called circular arrays, essentially all choices of distribution are
continuous so that the array element excitations are samples of the continuous
distribution. Tapered sidelobe space factors, including uniform excitation, will be
discussed later.
For applications requiring Dolph-Chebyshev-type sidelobes the circular Taylor h
distribution is used. This distribution is analogous to the linear Taylor h distribution
which is described in detail in Section 9.1.1. Only rotationally symmetric distribu-
tions are considered; the aperture distribution is a function of aperture radius p
only. The start is the uniform excitation pattern 2Ji(nu)/iru, where u = (D/X) sin 6.
D is the aperture diameter; all other symbols are as defined in Chapter 9. To pro-
duce h roughly equal-level close-in sidelobes at a specified level, h zeros on each side
of the main beam are modified. This is accomplished by removing (dividing) the
previous zeros and replacing them by n new zeros (multiplying). As before, a dilation
factor o is used to match the zeros at the h transition point. A single parameter A is
again used to control the sidelobe level. The sidelobe ratio:
SLR = COST17L4 (10.48)
Planar arrays 157

The space factor is given by:

nu n=i 1 — u 2ln%

The new zeros are:


un = ± a>/42 + ( « - | ) 2 , 1< w < h (10.50)
and the dilation factor is:

(10.51)

Zeros of J\(jtu) are jun. Beamwidth is the ideal beamwidth (which is the same as for
a line source, see Section 9.1.1) times o. Table 10.1 gives pertinent parameters for
various SLR and h. Since the space factors are much like those for line sources they
are not graphed here. The aperture distribution becomes (Taylor90):

HP) - I I ffig$ 00-52)


V m=0 \J\W)\
The radial variable p = 2np/d, and the coefficients Fm are:

ft (>-4
'ft,(.-£;
= 1,F M = -

Extensive tables of these distributions have been published by Hansen 4 2 ' 4 3


Aperture excitation efficiency is computed from:

2 (jog(p)pdp\
7? = — — (10.52)

When eqn. 10.52 is inserted and the integration performed, the terms in the
numerator for m > 0 contain the factor J0(7rjLim), which is zero, leaving only the
m = 0 term. In the denominator the integration of the product of the two series
yields an integral of two Bessel functions which are orthogonal, thus reducing the
product to a single series. The final result is:

n= -J F2 (10.55)
n
I +y
!

Table 10.1 Taylor circular n source characteristics


SLR
dB A u3 n=3 4 5 6 7 8 9 10
0

20 0.9528 0.4465 1.2104 1.1692 1.1398 1.1186 1.1028 1.0906 1.0810 1.0732
25 1.1366 0.4890 1.1792 1.1525 1.1296 1.1118 1.0979 1.0870 1.0782 1.0708
30 1.3200 0.5284 1.1455 1.1338 1.1180 1.1039 1.0923 1.0827 1.0749 1.0683
35 1.5032 0.5653 1.1134 1.1050 1.0951 1.0859 1.0779 1.0711 1.0653
40 1.6865 0.6000 1.0916 1.0910 1.0854 1.0789 1.0726 1.0670 1.0620
Planar arrays 159

Rudduck et al.m have computed a limited range of values, shown in Table 10.2. As
is the case for the linear Taylor h distribution, there is a value of n that gives maxi-
mum directivity. From the Table it is seen to increase with sidelobe ratio.

Table 10.2 Excitation efficiencyvs. n


SLR h
dB 3 4 5 6 8 10
15 0.9497 0.8404 0.7166 0.6023 0.4254 0.3083
20 0.9820 0.9723 0.9356 0.8808 0.7506 0.6238
25 0.9151 0.9324 0.9404 0.9379 0.9064 0.8526
30 0.8377 0.8482 0.8623 0.8735 0.8838 0.8804
35 0.7708 0.7779 0.7880 0.8048 0.8153
40 0.7056 0.7063 0.7119 0.7252 0.7365

Hansen one-parameter distribution: A symmetric circular distribution analogous


to the Taylor one-parameter distribution for line sources was developed by
Hansen.46 The close-in zeros of the uniform pattern, 2JI(ITU)/TTU, are shifted to sup-
press the corresponding sidelobes. Calling the single parameter if, the space factor is:

F(u) = (10.56)

Similar to the Taylor one-parameter line source, this definition of the pattern is
valid for u>H. For u < H, the form is

F(u) = (10.57)

Jt and 11 are the usual Bessel and modified Bessel functions of the first kind and
order one. Analogously to the (sin TTU)ITTU case, the pattern form for u > H provides
a sidelobe structure much like that of 2Ji(nu)/TTUi and part of the main beam, while
the pattern form for u < / / provides a higher main beam peak than unity, thereby
raising the sidelobe ratio. The transition from Ix to Jx occurs at w = //, and from
this point to the first sidelobe is 17.57dB. From this transition point to the beam
peak is 2II(7TH)/ITH, SO the sidelobe ratio (see Chapter 9 for definition) is:
2/ X (7T//)
SLR = 17.57 dB 4- 20 log ; (10.58)
nH
From this the sidelobe ratio vs. parameter// can be obtained (see Table 10.3). In
this Table, values are given up to 50dB SLR, as the current antenna art allows
50 dB to be realised. Since both Jx and/ x are represented by absolutely convergent
series and are quickly generated by common computer subroutines, the one-
parameter pattern is easy to use.
The aperture distribution is:
g(p) = (10.59)
160 Planar arrays
Table 10.3 Characteristics of Hansen one-parameter distribution
SLR Edge taper
dB H rad V dB
17.57 0 0.5145 1 0.5145 0
20 0.4872 0.5393 0.9786 0.5278 4.5
25 0.8899 0.5869 0.8711 0.5113 12.4
30 1.1977 0.6304 0.7595 0.4788 19.3
35 1.4708 0.6701 0.6683 0.4478 25.8
40 1.7254 0.7070 0.5964 0.4216 32.0
45 1.9681 0.7413 0.5390 0.3996 38.0
50 2.2026 0.7737 0.4923 0.3809 43.9

Fig. 10.13 shows these radial distributions for sidelobe ratios of 20(5) 40 dB. The
3 dB value is a solution of the following equation and is also given in the Table:

(10.60)
y/2H H2

0 .1 .2 .5 .6 .9 1.0
Aperture Radius

Fig. 10.13 One-parameter circular distribution with sidelobe ratio as parameter

Beamwidth is twice the u3 value. Excitation efficiency results from integration over
the aperture:

2(Jo g(p)pdp
V = (10.61)
Planar arrays 161

The Table also gives the product of beamwidth u3 and efficiency. In comparison
with the Taylor h circular distribution, the Hansen distribution is slightly less
efficient, but the required distribution is smoother and more robust.

Element pattern: When dipole or slots are used in planar arrays the spherical
coordinates as usually used do not allow the isolated element pattern to be written
simply, as in Chapter 9. In spherical coordinates, where the dipole is along the
jc-axis, the dipole pattern is:
IT
cos I — sin 9 cos 0
(10.62)
cos2
When 0 = 0 or 0 = TT/2, the result is unity. In the x-y plane (0 = TT/2) the pattern is:

h
cos I — cos <
F = ^ (10.63)
sin 0
while in thex-z plane (0 = 0) it becomes:
fa \
cos I — sin 8\
l ;
F = (10.64)
cos 6
10.2 Array feeds

10.2.1 Fixed beam arrays


Planar arrays may be composed of linear arrays, or may be an assemblage of individ-
ual elements. When the main beam direction is fixed, the constituent branch linear
arrays and the main feed line that feeds the linear array can be travelling wave or
resonant, or corporate. Thus, the flat plane array (FPA), which has replaced
parabolic reflector antennas in many aircraft and missile radars, is comprised of
branch 'sticks' which are resonant arrays of longitudinal shunt slots in waveguide,
with the waveguide sticks fed by a resonant main feed. Fig. 10.14 shows a quadrant
of the FPA, where it can be seen that each stick array is divided into sub-arrays,
with five comprising the centre stick. This is done to allow increased bandwidth
since each resonant array now has fewer than 21 slots. The complete array has 1368
slots. Fig. 10.15 shows the back, where the cross-guide feeds can be seen. A
waveguide corporate feed excites the cross-guide feeds. Since each quadrant has its
own feed port, a hybrid monopulse feed network can be used to provide sum and
azimuth and elevation difference patterns. Fig. 10.16 shows a very low (— 50 dB)
sidelobe edge slot waveguide array shaped to fit into a radome. Here the stick arrays
are travelling-wave arrays; since the array rotates in azimuth the beam squint is
easily accommodated. Achievement of such low sidelobe levels requires meticulous
162 Planar arrays

accommodation of mutual coupling and edge effects, and, of course, precise


fabrication. Another unusual type of array is a planar array of waveguide elements,
with a slab line feed designed for displaced phase centre operation.

Fig. 10.14 Quadrant of flat plane array {courtesy: Antenna Dept, Radar Systems Group,
Hughes, El Segundo)

Fig. 10.15 FPA corporate feed (courtesy: Antenna Dept, Radar Systems Group, Hughes, El
Segundo)

Fig. 10.17 shows one of 32 columns of a planar array of waveguide elements,


probe fed, at 1 GHz. The air slab-line tapered power divider is in the upper picture.
Planar arrays 163

Fig. 10.16 4W/4CS /OM/ sidelobe array {courtesy: Westinghouse Defense & Electronic Systems
Center)

Fig. 10.17 Waveguide elements and slab line feed {courtesy: Tsandoufas)
Side view and front view of Z.-band radiating module designed for low-profile
high-precision airborne application. Integrated construction consists of ground-
plane flange, air slab-line power divider feeding six air-filled waveguide radiators
via a securely captivated probe exciter. Thin (0.020 in) dielectric membrane pro-
vides airtight environmental protection. Module measures 2.5 ft X 4.3 in X 3.4 in
and weighs 4.5 Ib. Loss is less than 0.35 dB from input connector to array face
164 Planar arrays

A power divider for the rows incorporating ring hybrids for sum and difference
patterns is shown in Fig. 10.18 (Tsandoulas92). Air slab line reduces loss compared
with conventional strip line. Polarisation diversity may be obtained by using non-
resonant slots cut into the outer conductor of large square coaxial line at 4- 45°, with
every other stick using — 45°. Coupling is controlled by slot length; feeding
alternate sticks properly allows the radiated polarizations to be varied (Wolfson and
Cho107). Perhaps the simplest type of flat plane array is the resonant grid. This
scheme is a modification of the Kraus61 end-fire array, and consists of connected
wire rectangles above a ground plane (Tiuri et aL91). Fig. 10.19 shows a wire-grid

F ig. 10.18 Azimu th power divider using ring hybrids {courtesy: Tsandoulas)

Fig. 10.19 Wire grid array {courtesy: Conti and Toth)

FPA in quadrant form; each rectangle is resonant, so each quadrant is a 2-D


resonant array (Conti and Toth17) radiating VP. Bandwidth is limited both by the
microstrip spacing and by the resonant array ;larger arrays have less cross-polarisation
Planar arrays 165

and lower bandwidth. A major advantage is that fabrication is inexpensive. Tapered


excitation can be accomplished by making the vertical conductors wider near the
centre.
The radiating linear arrays and the main feed need not be the same type. The
main feed can also be a corporate feed. This type of feed is usually used to feed
individual elements. Many microstrip patch and stripline slot arrays are of this type,
although both resonant and travelling-wave printed-circuit arrays have been built.
Fig. 10.20 shows a corporate-fed stripline slot array, where the slots are the thin
dark lines. Another array, with orthogonal polarisation is located beneath; the slots
are thin dark lines which radiate through wide gaps in the upper array.

Fig. 10.20 Dual stripline slot arrays (courtesy: Physical Sciences Lab, New Mexico State
Univ.)

An unusual type of planar array is the image element array, originated by von
Trentini.95 In this scheme an element such as a slot in a ground plane is operated
with a thin high-dielectric-constant sheet parallel to and spaced about half wave (in
air) from the ground plane. The high e sheet is highly reflective, especially if it is
electrically quarter-wave thick, so that a set of images of the slot are set up. The
image element then acts as an end-fire array, with an effective area or gain that can
be as high as 15 dB above the slot alone. Thus these image elements can be used in a
widely spaced array with the narrowed element pattern suppressing grating lobes.
The ground plane and dielectric sheet act as a spatial filter, which is angle sensitive,
thereby producing low sidelobes from a sparse array. However at the Brewster angle
166 Planar arrays

(in the E-plane) the sheet is nearly transparent. The design must ensure that the
array has no grating or quantisation lobes near this angle, as they will not be sup-
pressed by the image element. When used in an array divided into quadrants for
monopulse tracking, the image element can provide improved difference pattern
sidelobes. Immell and Sasser49 found that two elements per quadrant gave
performance superior to a filled array of 19 slots per quadrant. In practice, the
elements can be waveguide or stripline slots or dipoles. The dielectric sheet may
have a relative permittivity of 30 or more; Ni-Al titanate is more machinable than
loaded resins. A honeycomb layer is used to support the dielectric sheet from the
ground plane.
Another unusual planar array is a flat plane array using a radial line feed.
Resonant slots are cut into the top plate of a radial transmission line; both radial
and circumferential slots (Goebels and Kelly32) and inclined slots (Kelly and
Goebels55), located in concentric rings, are used. However, the rectangular wave-
guide flat plane array allows more precise control of array distribution and
impedance.

10.2.2 Electronic scanning in one plane


When variable phase must be provided in one or both planes to scan the main beam,
the number of feed options is reduced. Phase shift may be provided by varying
phase through a phaser, or by varying frequency. Most electronically scanned
antennas operate in only one plane, with the antenna rotated to provide scan in the
other plane. Although each constituent-branch linear array could be equipped with
phasers, it is much simpler to put the phasers in the main feed. When a phaser is
located between branch-array coupling points, the main feed is termed 'series'. Each
phaser is set to the same value, and energy for the last stick goes in series through
all the phasers. The phasers can be ferrite or diode in waveguide, or in microstrip or
stripline (see Ince and Temme,50 for a review of phasers). Or the phaser can be a
length of waveguide or transmission line with appropriate frequency change. Such
series feeds must be travelling wave as there must be energy propagating in only
one direction in the feeder. A disadvantage of series feeds is that the first must
handle all the power, while the energy at the last stick experiences loss through all
the phasers. However, the equal phaser settings is an advantage. Fig. 10.21 shows a
planar waveguide array of edge slots, with baffles to reduce cross polarisation, with
ferrite phasers providing azimuth scan. 36 stick arrays are used, with 13 slots per
half stick. Frequency scanning in elevation via a snake feed is shown in Fig. 10.22.
The waveguide serpentine feeds 76 waveguide stick arrays, where each stick has 80
edge slots. Both main feed and sticks are travelling-wave arrays. Dielectric covers
make the slots invisible.
Parallel feeding occurs when each phaser is located at the front of each branch
array, after the main feed coupling. Now each phaser only handles 1/iVth of the
power but the phasers must be set to progressive multiples of the interelement
phase shift. Each branch array must be travelling-wave type. The ultimate parallel
feed is, of course, the corporate feed. Here a corporate feed, with phasers to the
Planar arrays 167

end of the 'fingers', can be used to feed the branch arrays. However, such feeds are
not compact.

. * • » „ - ; . . . , : ; : - : ; !

V, ' , ^ ' ^ . , ' v , .

Fig. 10.21 Ferrite scanned edge slot array {courtesy: Antenna Dept, Radar Systems Group,
Hughes, El Segundo)

10.2.3 Electronic scanning both planes


For scanning in two planes, two sets of phase shifts must be supplied. There are
only a few topological ways of achieving this. Fig. 10.23 shows the series-series
configuration. Here all azimuth phases are the same and all elevation phases are the
same. Both branch arrays and the main feed must be travelling wave. Impedance
mismatches in either branch or main arrays can cause extraneous beams (see
Section 9.2). The series-series configuration is seldom used for waveguide arrays as
there are no satisfactory ways of producing phase shift between waveguide slots in a
stick array. Ferrite posts or inserts have proved lossy and difficult to impedance
match.
A series-parallel configuration uses a series main feed line with phasers of equal
settings, and branch arrays where the phase rs are between the element and the
branch line coupling (see Fig. 10.24). This configuration is less common since the
branch arrays are not compact. The main line must be travelling wave and can be a
snake for frequency scanning. Branch lines can be either TW or resonant. Such
arrays can be implemented in stripline or microstrip, where each element is fed by a
phaser, with each row of phasers coupled to a TW or resonant feed. The advantage
is that these feeds are more compact than corporate feeds. However, the latter is
more forgiving of impedance mismatch.
The parallel-series configuration of Fig. 10.25 is the reverse of the previous one.
Now the branch arrays must be travelling wave while the main feed can be either
168 Planar arrays

TW or resonant. Again such a configuration is more suited to stripline or microstrip


than to waveguide.
With all the series feeds except the snake, there is the problem of integrating the
phaser into the feed. To allow a compact and inexpensive array, the phaser should
fit between the elements without folding or offsets. Most phasers are too large to

Fig. 10.22 Frequency scanned edge slot array {courtesy: ITT Gilfillan)

allow this. There have been integrated phaser-elements developed where ferrite
posts or diodes are used to change the phase of the waveguide mode. Those schemes
do not in general allow adequate control of mode amplitude and input impedance.
The parallel-parallel feed is corporate in that each element is connected to a
phaser, and each phaser is connected to the output port by an equal length path
Planar arrays 169

(see Fig. 10.26). A printed-circuit spiral array is depicted in Fig. 10.27, where
successive layers contain the baluns and power dividers, PIN diode phasers, and
corporate feed. The parallel-parallel configuration is often implemented as a distri-
buted array where each element is part of a receive-transmit module. One of the
early larger distributed arrays was the FPS-85;(see Reed80). Fig. 10.28 shows an
X-band array of open waveguides fed by a post located on the end of the module.

radiator

phaser

Fig. 10.23 Series-series feed

ne
radiator phaser

s—v
29 ^

Fig. 10.24 Series-parallel feed

The RASSR array and a module (Collins and Hayes;16 Opp et al.ls) are shown in
Fig. 10.29, This alumina microstrip module contains TR, receiver preamplifier,
transmitter power amplifier and diode phasers. Fig. 10.30 shows the EAR array and
a module; the array contains 1818 modules, each of \ in diameter. Wrapped around
the circular waveguide element and ferrite phaser is the phaser driver circuit.
170 Planar arrays

10.3 Mutual coupling

In this Section mutual coupling for both linear and planar arrays is discussed, as
there are many similarities between the two. After some basic material is covered,
element-by-element approaches are taken up. Following this, the periodic structure
approach is explained. Finally compensation for mutual coupling effects is discussed.

Fig. 10.25 Parallel-series feed

radiator phaser

Fig. 10.26 Parallel-parallel {corporate) feed


Planar arrays 171

ront

Fig. 10.27 Printed-circuit spiral array [courtesy: Raytheon Co.)

Fig. 10.28 Distributed waveguide array {courtesy: Antenna Dept, Radar Systems Group,
Hughes, El Segundo)
172 Planar arrays
10.3.1 Fundamentals

103.1.1 Current sheet: The simplest concept of a phased array is an infinite


flat current sheet carrying a uniform current flow parallel to one of the coordinate

Fig. 10.29A RASSR array {courtesy: Texas Instruments). B Element module {courtesy:
Texas Instruments)

axes, and is due to Wheeler.101'103 This current is phased to radiate at some desired
angle away from broadside. Such a current sheet can be used as a gedanken to
derive key scanning properties of phased arrays. The sheet may be either receiving
Planar arrays 173

or transmitting with the two situations giving related behaviour. An incident plane
wave (receiving) case will be considered first. Fig. 10.31 shows the current sheet
in side view, for both a wave incident in the 2^-plane, and in the //-plane. If the

Fig. 10.30A EAR array {courtesy: Westinghouse Defense & Electronic Systems Center)
B Element module {courtesy: Westinghouse Defense & Electronic Systems Centre)

current sheet is matched for normal incidence (6 — 0) to 77 ohms per square, then at
other scan angles there is a mismatch, as follows. For //-plane incidence (scan) the
incoming wave 'sees' a section of current sheet which is wider than the section of
174 Planar arrays

wavefront; thus the apparent resistance is lower by cos 6. The reflection coefficient
is now:
cos 0 - 1 - 0
V = —— = - t a n 2 - - (10.65)
cos 9 + 1 2
^-plane incidence gives the opposite result; the incoming wave sees a section of
current sheet which is longer than the section of wavefront, and thus the apparent
resistance is higher by 1/cos d. And the reflection coefficient is:
sec B - 1
r
Q
= taniz — (10.66)
sec 6 + 1

1/ H
\T /

(
/
—•H /
/ OH /
^E

(a) Ay-plane Kb) E-pIane


Fig. 10.31 Electric current sheet

H-plane
R
Ro

E-plane
90

Fig. 10.32 Current sheet active resistance {after Wheeler)

When the current sheet transmits instead of receives, the reflection coefficient signs
will be reversed. Fig. 10.32 shows the variation of current-sheet resistance forE- and
//-plane scans, for transmitting, with Ro = 17. Of course, this simple current sheet
model gives no information about reactance. The electric current sheet requires an
open-circuit reflector behind the sheet, where open circuit means c = 0, 77 = °°.
Magnetic current sheets may also be considered, with E and H interchanged. Now
the reflector behind the sheet is a short circuit a = °°. Thus the magnetic current
sheet, unlike the electric current sheet, may be approximated physically. An array
of short slots in a metal sheet, for example, provides such an approximation. From
Planar arrays 175

the foregoing discussion and Fig. 10.32 is indicated the basic behaviour of phased
arrays, where active resistance increases in one scan plane, but decreases in the
other. This trend will be observed in later Sections where precise active impedances
or admittances are calculated.
Further insight can be obtained from another concept due to Wheeler;103
namely, that of an ideal element pattern. Consider again a transmitting current
sheet, but let the current sheet be composed of short dipoles with patterns:
i/-plane F(d) = 1
£-plane F(0) = cos B
The patterns are affected by the effective (projected) aperture which varies as cos 0,
where 6 is measured from the normal to the array. Since the effective aperture
broadens the beam, the radiation resistance is proportional to sec #, so that the net
effect is:
R F2
(ia67)
F, = ^
For the two planes this gives:
JD
//-plane — = sec 6
Ro
R
£-plane — = cos B (10.68)
Ro

in agreement with the current-sheet values. If the element power pattern were
cos B, the resistive part of the array impedance would be matched. Thus the 'ideal'
element pattern is F(0) — cos1/2B. This is a conical pattern, symmetric about the
axis. The ideal element pattern has been approximated by a Huygen's source
which is a crossed electric dipole and magnetic dipole, sometimes realised by a
dipole and loop. However, the pattern of the Huygen's source is:

F(6) = §(l + cos#) = c o s 2 - (10.69)

and this is only a fair approximation to cos 1/2 #. In the Section on scan compen-
sation it will appear that both electric and magnetic modes are needed.* The
Huygen's source is then a crude approximation, giving a single electric and a single
magnetic mode.

10.3.1.2 Free and forced excitation: Arrays may be analysed from two concep-
tual element-by-element viewpoints. In the forced excitation model a driving voltage

* An alternative technique uses a medium over the array face with appropriately different
parallel and perpendicular transmission coefficients.
7 76 Planar arrays

(current) is applied to each element with the excitations phased from element-to-
element to produce the desired beam-scan angle. Each element has an active reflec-
tion coefficient which varies with scan angles, and which affects the actual gain
realised. The active reflection coefficient, of course, is associated with an active
element impedance (admittance). In the forced excitation model the active element
impedance (admittance) is calculated, from which the currents (voltages) at each
element may be obtained. For a dipole array the element currents are a solution of
the following equation, where the impedance matrix contains all of the inter-
element mutual impedancesZ^. The matrix is symmetric Z^ = Z^:
[V] = [Z][I] (10.70)

The overall array pattern is given by the sum over the array elements with the
currents as coefficients, all multiplied by the isolated element pattern. From this
the active element pattern can be obtained by factoring out the array factor. It
appears that this type of analysis is relatively easy to carry out, and this is true.
However, implementation of such an array is difficult as each element must be
fed by a constant-voltage (current) source. Simple feed networks in contrast are of
the constant-available-power type where an element impedance mismatch reduces
the applied voltage.
Free excitation assumes that each area of the feed network is equivalent to a
voltage (current) source in series with Ro. These are thus constant available power
sources. Such sources, with constant incident power, are suited to a scattering
analysis:
IK] = [S][Vt] (10.71)

Here Vt and Vr are the incident and reflected voltage (current) vectors and S is the
scattering matrix of coupling coefficients. The active reflection coefficient is given
by:

Ta = I I O ^ T ^ (10.72)
P Q

where C is the coupling coefficient between the 'oo' element and the 'pq* element.
The excitation coefficients differ by the progressive scan phase between the two
elements, and may also differ in amplitude if a finite array with a tapered excit-
ation is used. Although the scattering approach accurately represents most arrays,
and is conceptually simple, there is no direct way of calculating the coupling
coefficients. For isolated elements in a ground plane, the coupling in the //-plane
decreases as 1/r2 while that in theiS'-plane decreases only as 1/r (Wheeler102). In an
array, however, the coupling decay becomes asymptotic to 1/r2 as shown by
Hannan,39 Galindo and Wu29 and Steyskal.89 This is borne out by measurements on
large arrays (Amitay et a/.;3 Debski and Hannan18). Phase measurements show that
the coupled energy has the phase velocity of free space provided there is no external
loading.
Planar arrays 177

In some arrays there is a beam position (other than 6 — 90°) where the active
element pattern has a zero, or in different terms where the reflection coefficient
has magnitude of unity. Such an angle is called a 'blind spot'. It will be shown
later that the appearance of these can be precluded by proper choice of array
parameters.

103A.3 Active impedance and active element pattern'. In studying the


behaviour of a phased array the variation of gain and input impedance (at each
element) with scan angle is of critical importance. Either of two representations can
be used for the gain. First the array gain (pattern) can be decomposed into the pro-
duct of an array factor of isotropic, uncoupled elements and an active element gain
(pattern); the effects of mutual coupling are subsumed into the active element
pattern. Alternatively the array gain is decomposed into the product of array factor,

V V V V V

V V V V V

Kj \QA Kj m K)

Fig. 10.33 Two views of array gain


(a) Using active element pattern
(b) Using active impedance

isolated element gain pattern, and a factor displaying the impedance effects
explicitly. The active element pattern is measured with that element excited and all
other array elements present and terminated in Zg. In contrast, the isolated element
pattern is measured with all other elements open-circuited. This is not quite the
same as with all other elements absent, except for canonical minimum scattering
antennas (see next Section), see Fig. 10.33.
178 Planar arrays

For the active element pattern approach, call the array factorNh(6), where th<
peak value of h is unity. Only one space variable is shown for simplicity and N ii
the number of elements. By definition the field of the array is produced by the
active element times the array factor. Similarly the array power is N times the
active element power. Using the definition of gain, the array gain is:

= Nga{0)h\6) (10.73)
MR

where ga is the active element pattern. The formula and concept are simple, but the
measurement, with many terminated elements, is difficult.
Exactly equivalent is the open-circuited element pattern approach. The array
field is written as:
E(6) = NIaEoc(d)h(d) (10.74)
where Ia and IQC are the active and open-circuited driving currents. The open-
circuited element gain is immediately:

Unit voltage has been used for convenience. The open-circuited element resistance
is Roc. Now the array gain, with a matched load at angle 0, is:

ocv K J
G(9) = ' ° " (10.76)

Ia is again the active current and Rg is the generator resistance, which allows the
array gain to be written in terms of the open-circuited element gain:
4NRiRggoc(d)h\6) „„„
_j_ 7 |2 ^ '

A conjugate matched load reflection coefficient is defined as:

(10.78)

Note that this is the normal reflection coefficient with the Za in the numerator
conjugated. The active array gain then becomes (Allen2):
MR-
G(6) = —±gi(d)h2(d)\i-r\2 (10.79)

Since Ra\ 1 — F\2/Rg = 1 — I F | 2 , the array gain is also:


Planar arrays 179

Now the two expressions for array gain can be equated, resulting in a relationship
between active and open-circuited element gain:

Note that Ra and T are calculated at the angle 0; the equation relates active element
pattern to element active impedance (Oliner and Malech76).
It is interesting to examine the formula in detail at broadside with a match there.
Then the product of element directivity and input resistance for a half-wave dipole
above a ground plane is (Kraus60):
gtRi = 480sin2£/z (10.81)
76
and the active resistance at broadside is (Oliner and Malech ):

*o(0) = -^-sin2 kh (10.82)

Then the active element gain is ga = 4itDxDy. For half-wave spacing the array gain
is:

Garray = 4nNxNyDxDy = -^ (10.83)

where A is the array area. Since this 'array' is a part of an infinite array (uniform
excitation) the result is as expected.

10.3.1 A Isolated, open-circuited, and CMS elements: In understanding the


relationship between actual and ideal array element impedances and patterns, the
concepts of minimum scattering antennas (MSA) and canonical minimum scattering
antenna (CMS) are useful (Montgomery et at.;70 Kahn and Kurss;54 Gately etal.;31
Wasylkiwskyj and Kahn96). Consider an antenna withN ports; for example, these
ports might be the dominant mode and higher (usually evanescent) modes in an
open-ended waveguide element. When each port of an MSA is terminated in the
proper reactance, the scattered power Sti is zero. The CMS antenna is a special
type of MSA in that an open circuit at each port produces Su = 0. The CMS
antenna is lossless, which implies that the scattering matrix is unitary. The
scattering matrix can be written* as:

00 84)
-
This Af-port antenna possesses N orthogonal radiation patterns, which are 5,-y. In
the absence of non-reciprocal components, e.g. ferrites, the antenna is reciprocal,
which makes all patterns real, and symmetric about the origin; i.e. reversal of a
!
Note that + denotes the conjugate transpose
180 Planar arrays

line to any point on the three-dimensional pattern gives an equal point on the
pattern. Scattered and radiated patterns are equal. No impedance or admittance
matrix exists, and the N eigenvalues of the scattering matrix are all — 1. Unlike
most scatterers (or antennas) the scattered field pattern of a CMS antenna is
independent of the incident-field type and direction, although the amplitude of
the pattern will depend upon the incident wave. If, then, the pattern of an element
is taken in the array environment with all other elements open-circuited, this
pattern is the same as the isolated array pattern only if the element is a CMS
antenna. For example, when dipoles in an array are open-circuited, there is a
small effect due to the half-length conducting rods that remain when each
dipole is open-circuited. As expected, half-wave dipoles are not CMS antennas.
Similarly when slots in a ground plane are shorted across the feed terminals
the remaining half-length slots affect the current distributions on the ground
plane.
Wasylkiwskyj and Kahn96 showed that the mutual impedance (admittance)
between two identical MSA can be written as an integral of the power pattern
of an isolated element over certain real and complex angles:

Z 12 = 2 J p a t h exp (r-jkd cos w)P(w) dw (10.85)

where w = u +jv, and the power pattern is defined as P(w) =f(w)f(w + n) and
the integration path in the complex w plane is from (— 7r/2, — °°) to (— 7r/2, 0) to
(3/2, 0) to (3/2, + °°). The power pattern is normalised so that its integration over
all real angles is unity. This integral can be expressed in wavenumber coordinates.
For example, with two antennas in thexy plane, the mutual impedance is given by:

Mutual coupling, then, is specified by the element pattern and the lattice, and is
completely independent of the means utilised to produce that pattern. This
development for MSA is a generalisation of that of Borgiotti which is discussed in
Section 10.3.2.
Most multi-mode elements do not have orthogonal patterns, and hence are not
MSAs. However, many single-mode antennas approximate to an MSA. Short dipoles,
where the current distribution is essentially linear, are closely MSA. A resonant
(near half-wave) dipole is approximately MSA if the radius is very thin; this thinness
forces the current to be nearly sinusoidal. Mutual impedance between thin collinear
half-wave dipoles is shown to agree with the Carter results (Wasylkiwskyj and
Kahn96). Andersen, Lessow and Schjaer-Jacobsen5 show that crossed dipoles are
approximately MSA in one pattern plane, and that the mutual impedance calcu-
lated from MSA theory is good if the centre of one cross lies on a line bisecting the
arms of the other cross. Small helices are also approximately MSA.
Planar arrays 181

10.3.2 Element-by-element approaches


This Section approaches the array as a finite assembly of radiating elements. In con-
trast, Section 10.3.3 treats the infinite array. For finite arrays there are a finite
number of variables which allow certain types of analyses to be used. Basically an
impedance (or admittance) matrix relates applied voltages and the resulting
currents. Before proceeding to this analysis, however, mutual impedance between
only two elements is considered. Following the impedance matrix solution the
grating lobe series, active element efficiency and matching networks are discussed.

10.3.2.1 Mutual coupling between two elements: There are several intuitive
rules that concern mutual coupling between two antenna elements. First, the mag-
nitude of mutual impedance decreases with distance. Secondly, if one antenna is in
the pattern maximum of the other, coupling will be strong compared to that which
the pattern null produces. If the antenna is in the radiating near field of the other,
then the approximate near-field pattern is used. Thirdly, if the field polarisations
are parallel, coupling will be stronger than if they are collinear. For wire antennas
this can be restated in terms of shadowing: large shadowing will correlate with large
coupling. Fourthly, larger antennas have smaller coupling. For example, large horns
have lower mutual impedance than small horns. These rules, although useful, are no
substitute for actual values of mutual impedance or admittance. Since most arrays
use dipoles, slots, patches or horns, the discussion will be limited to them.
In element-by-element approaches it is necessary to start with mutual impedance
(or admittance) between elements. Since these are often slots or dipoles, and since
slots can be replaced by Babinet flat-strip dipoles, which if narrow in turn can be
replaced by cylindrical dipoles, the mutual impedance between dipoles is a basic
building block. For mutual impedance the zero-order (sinusoidal current distri-
bution) theory is usually adequate. For this case, the formulas of Howard King56
based on the original work of Carter14 can be used. They give mutual impedance
between parallel coplanar dipoles of unequal length in echelon. However, these
formulas contain 24 sine and cosine integral terms, of various complicated and
diverse arguments. Although the results of King could be rearranged and grouped
into a computer algorithm, it is simpler to directly "derive the formula, using the
exact near electric-field formulations of Schelkunoff and Friis.85 They show that
the exact field from a sinusoidal current is written as three spherical waves, although
the field components are in cylindrical coordinates. Mutual-impedance is then the
integral of this electric field (from each half dipole) times the current distribution
at the other dipole. With the geometry of Fig. 10.34 the mutual-impedance
expression is:
— ;30
Z = I [to - 2^ 2 cos kd2 + to + to - 2to cos kd2 + i//6]

x sin A : ^ — x)dx
where
ex
k = 2TT/X, Sx = sin kdu S2 = sin kd2, to = P( # * * ) (10.87)
182 Planar arrays

Thei? values are:

(10.88)

The integral could be evaluated numerically, e.g. by Romberg integration, and for
large slot separations this method is preferable. However, the adjacent and nearby
slots are most important, and for these an exact evaluation is necessary.

" 2 d !T

Fig. 10.34 Echelon dipoles

The exact solution can be written as two sums (Hansen and Brunner47), where
n steps by 2:

1 5
" ~ I I Cm[ejku[E(kA-kU)-E(kB
m= -l n = -

+ e~iku[E(kA -kU)-E(kB-kV)]] (10.89)

The coefficients are:

d = Cx = 1, Co = -2coskd2

and^(x) is the exponential integral (Abramowitz and Stegun1) with arguments:

U = dx+ n(xQ + md2)


V = n(xo + md2)
A2 = y20 + (x 0 + nd2 + md2f
B2 = yl + (xo + md2)2 (10.90)
Planar arrays 183

When the two dipoles are the same length the formula simplifies (Hansen45) to:

Z = M mt= - 2 n I= - l Am
«J

where
j3 = \/(x0 + mdf +yl-n(xo + md) (10.91)
and the coefficients are:
A_2 = A2 = 1, v4.j = Ax = — 4cosfoi
2
,4 0 = 2(1 4- 2 cos kd) (10.92)
When j o = 0, i.e. the dipoles are collinear, y0 is replaced by the dipole radius a.
When the dipoles are half wave, the m = 2 and 4 terms disappear and the exponen-
tial simplifies. This formulation is similar to that of Richmond83 for two equal-length
thin dipoles with axes at an angle. Computer subroutines for Ci, Si are readily
available in most computer libraries, with computation time comparable to that of
trig functions. These subroutines use the economised series developed by Wimp and
Luke.106
Fig. 10.35 shows Zy between two parallel half-wave dipoles o n a i ? +jX plot.
The curve, with spacing/wavelength as a parameter, is similar to a Cornu spiral. It
can be seen that the magnitude of impedance decreases as spacing increases. Data
are normalised by the self resistance, which, for a zero-order (zero-thickness)
dipole, is 73.13 O. Fig. 10.36 is a similar plot for collinear dipoles. The spacing here
is between dipole tips; for centre-to-centre spacing add 0.5. As expected, the
collinear coupling is less owing both to lower shadowing and a pattern null. Recall,
however, that the dipole near field has both axial and radial electric fields and that
the latter does not have a pattern null along the axis.
When slots are located in a stripline surface, or in contiguous waveguides, the
mutual admittance between two slots behaves as though the slots are located in an
infinite ground plane, to a very good approximation. The slots can then be replaced
by strip dipoles by Babinef s principle developed by Booker10 for self impedance,
and also for mutual impedance (Begovich9). The slot mutual impedance Z 12 and
strip dipole mutual admittance. Yn are related:
2Z 12 = r?2r12 (10.93)
where 17 = 12O7T. The factor of 2 has been reduced from 4 as the slots radiate one
side only.
When the slots are narrow, which is usually the case, the equivalent strip dipoles
are also narrow; they can be accurately replaced by cylindrical dipoles of radius
equal to one-fourth the slot or strip dipole width (Lo 65 ). Thus the extensive
development of dipole mutual impedance can be used. As discussed in Section 9.2,
the Carter dipole mutual impedance is satisfactory for the design of linear slot
arrays.
184 Planar arrays

Mutual impedance between microstrip patch elements has not received much
attention, probably because arrays of these most often employ a fixed beam. In
principle, an approximation to mutual impedance could be obtained for rectangular

0.05

-0.4 -0.2
R/R

Fig. 10.35 Parallel dipole impedance


Spacing at 0.05X intervals

01r

005

R/Rn

Fig. 10.36 Collinear dipole impedance


tip spacing at 0.05X intervals

patch elements by assuming that two opposite gaps act as virtual radiators for linear
polarisation (see Chapter 9). Then a Carter type of zero-order theory could be used
Planar arrays 185

to find the mutual impedance to the TEM mode inside the patch, and this
impedance could then be transformed to the feed terminal. More sophisticated
techniques such as moment method are probably not necessary as the field distribu-
tion across the patch gaps is approximately the canonical value. A recent approach
by Krowne and Sindoris62 calculates //-plane coupling between rectangular patch
elements by replacing the two radiating gaps by two coupled microstrip lines, where
the first is excited and the third acts as receiver. Some measured data are given by
Jedlicka and Carver.51

-AOr

-50 -

-60 -

-70 -

- 80
-180 -90

deg

Fig. 10.37 Coupling between E-plane sectoral horns {courtesy: Lyon)

Coupling between horns has been investigated by Lyon et al.66, who find that
the coupling is sensitive to orientation, as the polarisation rule suggests. Fig. 10.37
shows coupling between identical £-plane sectoral horns with 8 dB gain, at a separ-
ation of 3.73X. As the polarisation rule predicts, the//-plane and/T-plane couplings
alternate peaks and valleys as the orientation angle is changed. For large spacing the
parallel polarisation decays as 1/r2, or 12dB for each doubling of distance r, while
186 Planar arrays

the collinear coupling decays as 1/r. Coupling data for pyramidal horns of a separ-
ation of 5.93X are shown in Fig. 10.38. Hamid36 used GTD to calculate mutual
coupling between sectoral horns with 3X apertures; the coupling showed a typical
peak and null behaviour with centre-to-centre separations, with the envelope vary-
ing roughly as coupling ~ (— 49 — s/X) dB.

-AOr

-50

-60
CO

en
c
"Q.

8 -70

-80
-180 -90 + 90 + 180

Fig. 10.38 Coupling between pyramidal horns (courtesy: Lyon)

10.3.2.2 Impedance matrix solution: Finite arrays of single-mode elements


such as dipoles or slots can be analysed simply by setting up a mutual-impedance
matrix. Carter mutual impedances (Section 10.3.3) are usually adequate. If TV
dipoles are used over a ground plane, it is replaced by an image array, which makes
the impedance matrix 2N x 2N instead of N x N. With a symmetric array the
matrix can be folded to reduce its size to roughly half (exactly half if N is even).
The scan angle appears in the progressive phase of the drive vector of which the
component from the nth row and mth column is:
Vt = Anm exp [—j2-n(nDx sin 60 cos 0 O sin 60 sin 0o)] (10.94)
The mutual impedance Ztj is between the zth element and the/th element. When
the complex simultaneous equations are solved the result is the current vector:
* Note that solving simultaneous equations is roughly three times faster than matrix inversion
(see Westlake100)
Planar arrays 187

(10.95)
and from the current vector the active impedance of each element in the array is
determined. When the solution is desired for several scan angles, so that only the
drive vector changes, but not the Zzy matrix, a simultaneous equation solver can be
used that stores a diagonalised matrix, allowing a rapid solution for new vectors
after the first. From the current vector (active-array excitations) of course array
patterns can be calculated. The impedance-matrix solution has an intrinsic advant-
age: array excitation tapers can be incorporated directly into the drive vector, unlike
the infinite array methods of Section 10.3.3 where the excitation must be uniform.
Figs. 10.39 and 10.40 show active resistance for the centre element of a 7 x 9 array
of half-wave dipoles, in free space, and quarter-wave over a ground plane. Also
shown in Fig. 10.40 is active resistance for comparison for a very large array. It can
be seen that small arrays add an oscillatory behaviour to the infinite array curves.
Active reactance for the centre dipole of an array over a ground plane is given in
Fig. 10.41, where again the oscillations can be observed. Active impedance of
dipoles near the edges behaves differently, but almost no data has been published.

200 r

80

Fig. 10.39 Active resistance of 7 X 9 array of half-wave dipoles with 0.5\ spacing (courtesy:
B. L. Diamond)

Multi-mode elements in a finite array can, in principle, be handled by increasing


the size of the matrix from N x N to 6N x 67V, where each element has six modes,
for example. Even for linear arrays this poses problems as the computer time for
solution of simultaneous equations varies as the cube of the matrix size. An
alternative solution has been offered by Goldberg.33 The mode voltage at each
element is written as a sum (over the elements) of drive voltages times coupling
coefficients times scan phase factor. The coupling coefficients are approximately
independent of array size (strictly true only for CMS elements) and may be
188 Planar arrays

200 r

^--^\ H-plane

0 20 40 60 80
6,deg

Fig. 10.40 Active resistance of dipole array over ground plane [courtesy: Diamond)
infinite array

40 50 80 100 120 140 160 180


Ra(9,0)

Fig. 10.41 Active reactance of arrays of half-wave dipoles with 0.5 X spacing [courtesy:
Diamond)
65X149

9 X11
Planar arrays 189
determined from an infinite array. An infinite array with the proper spacing, scan
range, and bandwidth is then designed as described elsewhere in this Section, with
the result of mode voltages and reflection coefficient vs. u. A Fourier inversion is
used to obtain the coupling coefficients, which are then summed to give a set of
mode voltages and reflection coefficient for each element in the array. Only the
centre elements in a large array experience essentially all the coupling; elements at
or near the edge employ fewer strong coupling coefficients as they decay rapidly
with element separation except near grating lobe incidence. Effects of amplitude
taper can be determined simply by including the amplitude coefficients in each
sum. From the reflection coefficient of each element the active impedance can
be written, and the pattern of the array is found by adding the modal array patterns
which use element modal patterns and mode voltages. This multi-mode finite
array solution has been extended to two-dimensional slot arrays by Woody and
Hansen.109
Active element patterns are of interest for small arrays although it is no longer
valid to multiply the array factor by a single active element pattern in order to get
the overall active array pattern. Fig. 10.42 shows calculated and measured active
element patterns of the centre dipole in a 5 x 5 array of half-wave dipoles in a half-
wave lattice, and quarter-wave over a ground plane. The broadening and flattening

: / *

Vw
/calculated - • measured \

A
10-
dB i
• E-plane
1 I 20-"
1 I
i i i I i
s 5
90* 60 3 0 * 0 * 30* 60 90°

90° 60° 30° 0° 30° 60° 90°

Fig. 10.42 Active element patterns of 5 X 5 array of dipoles over ground plane {courtesy:
Rupp)

effect on the E-plane pattern is quite noticeable. Figs. 10.43 and 10.44 are active
element patterns for the centre element of a 7 x 9 array of dipoles for various
lattice spacings. From these curves it is apparent that the active-element pattern
limits the scan-angle range. As more elements are added to the array the number
of oscillations in the pattern increases and the curves become smoother. Also
190 Planar arrays

the fall off becomes steeper. Figs. 10.45 and 10.46 compare centre and corner
active element patterns for the 9 x 1 1 array of dipoles previously discussed. The
corner elements do not show the symmetric mutual coupling effects that tend to
flatten the active-element pattern. Dummy elements can be added around the edges
of an array to be scanned if space is available. The general design procedure is to
design using infinite array parameters to obtain the basic array parameters, then
to refine this design to accommodate edge effects.

D x «D y -o.8X

Dx=Dy»0.7X

Fig. 10.43 H-plane active element pattern of centre element of 7 X 9 array of dipoles
{courtesy: Diamond)

isolated
dipole

-80 -60 -40 -20 0 20 40 60 80

Fig. 10.44 H-plane active element pattern of centre element of 1 X 9 array of dipoles over a
ground plane (courtesy: Diamond)

An interesting example of edge effects in a linear array is given by Gallegro,30


for half-wave dipoles with and without a ground plane. Spacing between elements
is 0.586X, which gives a just visible grating lobe at 6 = 90° for 45° scan; dipole
orientation is both parallel and collinear. The centre element is matched at 0°
Planar arrays 191
scan angle with all elements using the same matching network. All arrays have 51
elements, and the ground plane spacing is 0.25 X. The collinear arrays are well
behaved even at grating lobe incidence; the parallel arrays are somewhat less well
behaved up to GL incidence, but at incidence there was a marked edge effect,

-80 -60 -40 -20 0 20 40 60 80

Fig. 10.45 E-plane active element pattern of 9 X 11 dipole array over ground plane {courtesy:
Diamond)

Corner
element

Centre
element

-80-60-40-20 0 20 40 60 80
0,deg
Fig. 10.46 H-plane active element pattern of 9 X 11 dipole array overground plane {courtesy:
Diamond)

especially for the array without ground plane. Figs. 10.47 and 10.48 show element
impedance by number with and without ground plane for scan angles of 0°, 40°,
45° for parallel dipoles. It appears that the edge elements in the scan direction are
more affected than the edge elements in the reverse direction, but the VSWR of the
reverse edge elements is higher owing to the grating lobe pointing in that direction.
A rough tabulation of the number of edge elements where the edge element
impedance is outside the centre cluster is shown in Table 10.4. It must not be
supposed that these results apply to planar arrays, but the trends should be useful.
Another method of analysing finite arrays utilises correlation functions (Zaghloul
192 Planar arrays

and MacPhie,111). First, we have the spatial cross-correlation function, which is the
integral of the aperture field of the nth elemenl times the total aperture field. The
Fourier transform of this is the cross-power-density spectrum between the nth
element and the array. Integration of this over the u-v domain gives total power,
from which active element impedance is expressed as a sum of terms, one for each
element. These terms contain unknown element aperture fields times a scan
independent integral. The unknown element fields can be expressed in terms of
reflection coefficient or impedance, so that a set of simultaneous equations results.

45°

112

cj
X

scan
angle 0°

3,49
2,50

16
40 56 72 88 104 120

Fig. 10.47 Linear array element impedance, ground plane (after Gallegro)

This technique differs from the element-by-element approach using Carter


mutual-impedance terms in using a simpler scan-independent integral. The latter is
roughly a plane-wave mutual-impedance integral, whereas the Carter result is
Planar arrays 193

obtained from the spherical waves that comprise the exact near field. In terms of
computation and accuracy the correlation method does not appear advantageous.
The Carter method has an interpretive advantage in that all terms have a direct
physical meaning.

216 r

180

108

72

36
51
51

2L Z>8 72 .96 120 144


K ^1

Fig. 10.48 Linear array element impedance. No ground plane (after Gallegro)

10.3.2.3 Grating lobe series


It was observed by Wheeler104 that each term in the double series for slot active
admittance or for dipole active impedance is associated with a point in the u, v
grating lobe lattice, e.g. Fig. 10.49. This is true for all types of elements in a regular
lattice, but is most easily visualised by considering short slots or dipoles. For short
slots, the formulas for slots of general length simplify because the sine and cose
factors become unity. If now eqns. 10.117 and 10.121 (derived later in Section
10.3.3) are divided by the broadside active conductance, the normalised admittance
becomes:
194 Planar arrays

G«(«, v) _
(10.96)
Ga(p,o) ksjk2-klo-kl yo
Ba(u, v) __ _ k2—
(10.97)
Ga(p,o) , + ««-i

Table 10.4 Number of 'edge elements': grating lobe for 45° scan

Dipoles Ground plane Scan direction Reverse direction


parallel no 1-5-30 1-2-3
parallel yes 1-3-6 1-2-3
collinear no 1-2-2 1-1-1
collinear yes 1-1-2 1-1-2
Numbers are for scan angle of 0°-40°-45°

Z- crater

U] = A/Ux V] = A/Uy

Fig. 10.49 Grating lobe craters in u-v space

Here the expressions for H(oo and Hnm are identical with eqns. 10.118 and 10.123.
Wheeler correlated these terms with the u, v grating lobe points, and identified each
term as an impedance crater, a concept developed by Rhodes.82 The normalised
admittance becomes:
Planar arrays 195

n m
Ya(u,v)
(10.98)
Ga(o,o) Ga(o, o)
Here Ya/Ga is the admittance (impedance) crater which is a plot of the normalised
admittance (impedance) of an element in u, v space. Fig. 10.50 shows the crater for
a short dipole and Fig. 10.51 is the crater for a half-wave slot (or dipole). Wheeler
discovered that mutual-coupling effects with scan, and the effects associated with
GL onset, could be visualised by placing an impedance crater at each (inverse)
lattice point in the u, v plane. The term 'crater' arises because the resistance pre-
sents the topological form of a crater. Thus the grating lobe series is a link between
the current-sheet concepts discussed earlier and the periodic structure approach.

Fig. 10.50 Impedance crater for a small electric dipole, described by a contour map
{courtesy: Wheeler)
Resistance
Reactance

The grating lobe series can be written using open-circuited element patterns
instead of impedance craters. From the periodic cell calculations of impedance/
admittance in Section 10.3.3, the real part of active impedance is (Oliner and
Malech,;76 Wasylkiwskyj and Kahn96):
+Z 2
goc{6)\{ , • COS flgocffnm)]
(10.99)
dxdy g cos e [ cos0 nm ^ oc (0)J
196 Planar arrays

where the prime signifies omission of n = m = 0. This expresses the active resistance
as a sum of open-circuited element patterns evaluated at the grating-lobe lattice
points, for the visible grating lobes only. By using the relationship between the
active and open-circuited element pattern previously developed, the active element
pattern is written in terms of the sum of visible grating-lobe open-circuited element
patterns (Wasylkiwskyj and Kahn98):

(10.100)

cosdnmgoc(6)

ky
2k 3k 4k

/? / / / / / / / / / / / / / / / / /
3k
/ / / / / / /

////////////// rz

mmmmmm^^^^^^^^^^ kx

Fig. 10.51 Admittance crater of a half-wave slot (courtesy: Rhodes)

Comparing this result with that of Hannan,37 which is the same except that the
ratio of element patterns in the summation uses active patterns, thus the inference
is that an element pattern ratio as used is independent of the impedance termin-
ation of the other elements. That is, the ratio is the same for match-terminated and
for open-circuited element patterns. A further inference follows: that a matching
network may be connected between the elements and the generators without
changing these ratios. If the network allows T -> 0, then the active element pattern
will approach the ideal element pattern. For the case of no visible grating lobes, the
formula reduces as expected to:
Planar arrays 197

(10.101)
A"

As previously discussed the open-circuited element pattern is equal to the isolated


element pattern for minimum scattering antennas. Actual single-mode elements
such as thin dipoles and slots are approximately MSA.
The grating-lobe series can also be expressed in terms of transforms of aperture
fields (Borgiotti12). Active admittance of an infinite array on a rectangular lattice is
written as an integral in wavenumber space of the Fourier transform of the aperture
field times the FT of the conjugate field for each element. The doubly infinite sum
representing element phases can be expressed as a doubly periodic delta function,
which reduces the integrals to:

4TTX ^ ^ ,..„„*. „ „ „ (10.102)

where w = —jyu2 + v2—k2 and the fields are evaluated at the u, v grating-lobe
points. Mutual impedance between two elements can similarly be written as an
integral in wavenumber space of aperture distribution FTs, which can be reduced
to the aperture complex power equations developed by Borgiotti11 and Rhodes.81

10.3.2.4 Active-element efficiency: Efficiency of an element in an infinite


array is defined as the ratio of radiated to available power. Since radiated power
varies with scan angles, a precise definition is the average ratio (Kahn52):

T? = l-^j^\r\2dudv (10.103)

For linear arrays the corresponding definition is:

T? = 1 - - C\T\2du (10.104)
7T J O

Thus element efficiency is a measure of the effect of mutual coupling. But it is use-
ful in relation to the concept of 'ideal element'. For a square array lattice with
spacing of A/2 or less, no grating lobes exist. The ideal element power pattern, since
the ideal element has no reflected power for scan angles in visible space, is just cos 6.
However, there are array phasings that provide a beam outside the visible unit circle
inu,v space, and for these | F | = 1 since no radiation occurs. Thus the ideal element
in the non-grating-lobe case has mutual coupling, which changes F from zero to
unity as the scan vector cross the unit circle. If the lattice spacing is increased to
Vo3X so that radiation always occurs, as the scan vector crosses the unit circle, a
grating-lobe scan vector enters the unit circle. The ideal element pattern for the case
of grating lobes was first defined to be cos 6 (in power) out to the angle at which
the first grating lobe appears, and zero from that angle to 90° (Hannan37). A more
useful definition has been developed, based on the grating lobe series of the pre-
198 Planar arrays

vious Section. Now the ideal element pattern is the limit of the active element pat-
tern as 0 -> 0. That is, if a matching network were applied to the array terminals, as
the network allowed the match to approach a perfect match, the active element
pattern would approach the ideal element pattern. Recapitulating, with F = 0
(Wasylkiwskyj and Kahn99):
DxDy cos 0
iidedl(0) =
'cos0£ o c (0 n m )
1+1 (10.105)

where the prime indicates that the sum is over all visible grating lobes, but not
including the main beam (n = 0 = ra). For a linear array the ideal element pattern is:

Dx cos 0 (10.106)
8idedl(P) =
cos0g QC (0 n )
1
cosdngoc(d)
For example, consider a linear array of thin slots with spacing between X/2 and X.
The ideal pattern is:
gidedi(0) = Dx cos 0 for 0 < sin < l/Dx - 1
Dx cos 0
for 1/DX- 1 <sin0< 1 (10.107)
cos0
1 +•
COS0!
where cos2 0x = 1 — (sin 0 — l/Dx)2. Taking a specific example, let the spacing be
0.75X. Fig. 10.52 shows the ideal element pattern. Such a pattern would maximise
element efficiency.

0-8
0 7

06
^ 05
CD
~ 0u
Q.
0 3

0 2

0-1

0
0 0 1 0 2 0 - 3 0 - 4 0 - 5 0 6 0 - 7 0 - 8 0 - 9 1-0
sin 9

Fig. 10.52 Ideal element pattern (courtesy: Wasylkiwskyj and Kahn)

Applying eqn. 10.106 to half-wave dipoles yields the following results, where
angle 0 is zero at broadside.
Planar arrays 199

collinear, g ideal = 1 / 2

COS2 0
parallel, in plane of dipoles,
2 - sin2 0
1
parallel, in normal plane, gideal = (10.108)
2 - sin2 0
The collinear ideal pattern is isotropic, while the parallel in-plane pattern is roughly
like that of a half-wave dipole. Somewhat unusual is the behaviour of the parallel
normal plane pattern, which has twice the value at endfire. This is perhaps related
to the fact that apertures at endfire have twice the directivity as at broadside. Of
course, array designers attempt to avoid grating lobes; the utility of these ideal
element pattern concepts is in providing insight and understanding of the inter-
actions among element patterns, lattice and performance.
Returning to ideal element efficiency, eqn. 10.109 relates 17 to reflection
coefficient. In a space, where for a rectangular lattice of size a^ a2
(*! = 2nai sin 6 cos 0
<x2 = 2na2 sin 0 sin 0 (10.109)

1.0r

.2 0.5

0.5
element spacing,

Fig. 1 0 5 3 Ideal element efficiency for rectangular lattice {courtesy: Kahn)

Visible space is bounded by an ellipse in a space; the integral in eqn 10.103 is the
common area between the ellipse and the IT XTT cell in ex space. Fig. 10.53 shows
ideal element efficiency for an infinite array with rectangular lattice (Kahn52). Note
that for a square lattice with A/2 spacing, the ideal element efficiency is TT/4 as dis-
200 Planar arrays

cussed under the element-gain paradox. A regular hexagonal lattice (equiangular


triangle lattice) shows ideal element efficiency somewhat lower, in Fig. 10.54.
Finite arrays are less efficient owing to the edge coupling being different from
centre coupling. Wasylkiwskyj and Kahn97 calculate ideal element efficiency for a
square lattice. Fig. 10.55 shows this for broadside and end-fire scan. More than 10
elements per side are needed to approximate the infinite array efficiency reasonably
well. Thus an insight can be gained into the effects of reducing the size of an array.

1-0r

0-5

0-5 10
element spacing t a-j

Fig. 10.54 Ideal element efficiency for hexagonal lattice (courtesy: Kahn)

10.3.2.5 Impedance matching networks: Networks connected between the


antenna elements and the generators can be used either to make reflectionless
elements, or to make scatterless elements. The former is easier and will be discussed
first. By use of successively more complex networks, the F = 0 situation may be
closely approached (Hannan40). However an exact match at all angles can only be
approximated (Varon and Zysman93). It will appear later that a match may be
approached through use of an external network, i.e. a spatial filter network at the
array face. Both methods improve the match by making the active-element power
pattern closer to cos 6 (for a planar array with no grating lobes).
As a means of understanding the matching network process, consider the net-
work to consist of three networks in cascade. The first network is a 27V-port that
transforms the array excitation vectors into eigenvectors of the active impedance
matrix. The second network consists of TV separate 2-ports that equalise the eigen-
values. This process is analogous to the convergence speedup in adaptive arrays
where the vectors are first orthogonalised, then amplitudes are made equal. The
Planar arrays 201

third network is a 2N-port which is the inverse of the first; it restores the array
excitation vectors, which have now been matched. An excitation in an infinite
array producing a beam in invisible space is represented by a singular (zero or
infinite) eigen excitation. For finite arrays such excitations produce very small or
very large eigen excitations. Attempts to match such eigenvectors produce tolerance
problems similar to those encountered in superdirectivity. Kahn53 has calculated
active element patterns for a linear array of uniform line sources. Fig. 10.56a and
10.56Z? are for an element spacing of A/2, for the centre element of 15- and 25-
element arrays. The solid line is the ideal cos 6 and the dashed line is the actual
element pattern. Fig. 10.56c is for the edge element in the 25-element array. The
edge-element pattern is broadened and reduced slightly in gain. Fig. 10.57 is for
a 25-element array with 0.6366X spacing. Again the solid curves are the ideal
element pattern discussed elsewhere and the dashed curves are the actual element
patterns. Centre, edge, and next-to-edge element patterns are in the three graphs.

0 0-1 02 03 OX 0-5 06 0 7 08 0 0-1 0-2 0-3 OX 0-5 0-6 07 0-8


element spacing d/x element spacing. d/X

a End fire b Broadside


Fig. 10.55 Ideal element efficiency vs. number of elements {courtesy: Wasylkiwskyj and
Kahn)

M2 = 729
M2 = 121
M2 = 49
M2 = 9
M2 = 1

As expected the finite array size fills in the grating-lobe null somewhat. The
edge pattern is flatter, with reduced gain. In contrast, the next-to-edge pattern
has full gain but a rapid falloff. From these data there appear to be two to three
'edge' elements, which corroborates the experimental measurements of Gallegro
202 Planar arrays

(see Section 10.3.2.2). It was mentioned that invisible space excitations cause
extreme eigenvalues. For a 25-element array with spacing of 0.3183 X there are 17
eigenvalues of the order of unity, and eight that are very small. When the two
largest and all the very small ones are discarded, a usable match occurs, as evidenced
in the active element pattern of Fig. 10.58^. Use of all eigenvalues results in the
unstable pattern of Fig. 10.58&.

0 10 20 30 40 50 60 70 80 90 0 10 20 30 40 50 60 70 80 0 10 20 30 40 50 60 70 80 90
pattern angle.9 deg pattern angle.0 deg pattern angle 9. deg
b

Fig. 10.56 Active element pattern of linear array, d = 0.5 \ (courtesy: Kahn)
a N = 15 centre element
b N = 25 centre element
c N = 25 end element
cos 6 pattern

r grating lobe null at 34.3°

V
0 10 20 30 40 50 60 70 80 90 0 10 20 30 40 50 60 70 80 90 ' 0 10 20 30 40 50 60 70 80 90
pattern angle. 9 deg pattern angle. 6 deg pattern angle. 8 deg
b

Fig. 10.57 Active element pattern of linear array, d = 0.637 \, N = 25 {courtesy: Kahn)
a centre element
b next-to-end element
c end element
ideal element pattern

Descattering the array via a network requires that the element transmit and
receive patterns be equal, i.e. that the elements are minimum scattering (see
Section 10.3.1). In practice, these remarks are useful for single-mode elements such
as slots and dipoles. A further requirement is that all mutual impedances be reactive
Planar arrays 203

(Andersen and Rasmussen6). A corollary is that the mutual resistance, since it is


proportional to the beam orthogonality integral, must equal zero (Wasylkiwskyj
and Kahn96). Now the mutual impedance is related to the element power pattern;
so the zero mutual resistance becomes a constraint on the element pattern. For
example, rotationally symmetric power patterns of eos m 0 yield R12 = 0 for
spacings that satisfy:
jm/2(kd) = 0 (10.110)

where / m / 2 is the spherical Bessel function of order m/2. Pursuing the example
further for isotropic elements (m = 0) the root gives d = A/2, so that a half-wave
spaced array may be completely decoupled and descattered.

X>

j/> 1 0 10
o
Inor

5 0.5 S 05
Q. Q.
power

nn | ! 1 1 1 1 1 1 \(
w
0 10 20 30 40 50 60 70 80 90 0 10 20 30 40 50 60 70 80 90
pattern angle.Sdeg pattern angle, 8 deg
b

Fig. 10.58 Active element pattern of linear array, d — X, N = 25 (courtesy: Kahn)


a partial match
b complete match

When a grating lobe just enters visible space, the active element pattern has a null
(at that angle), or in active-impedance terms the reflection-coefficient magnitude is
unity. Such a blind spot cannot be removed by a matching network (Wasylkiwskyj
and Kahn"). However, for those blind spots that occur at scan ranges inside the
grating-lobe scan contour (in u-v space), a suitable network can improve or even
remove the null, provided that the open-circuited element pattern does not have a
null at that angle.

10.3.3 Periodic cell approach

10.3.3.1 Slots: The edge effects that make the analysis and understanding of a
finite array difficult are removed when the array is made infinite: the finite array is
204 Planar arrays

now considered as a portion of the infinite array. Because of Floquet's theorem,


the field at each element differs from that of adjacent elements by only a phase
shift. Using this concept, Wheeler101 deceloped the unit cell, or periodic cell
wherein each element is surrounded by a cell, with all cells alike and contiguous.
Edelberg and Oliner23'24 and Oliner and Malech76 extended the periodic cell
approach to scanned arrays. It has proved to be the most powerful and perceptive
technique for understanding and for designing sophisticated arrays. An equivalent
development is by Diamond.21 Because of the Floquet symmetry, the single unit
cell contains the complete admittance behaviour of each element in the array. The
unit cells are normal to the array face, centred about the elements, and contiguous;
see Fig. 10.59. For broadside radiation with no grating lobes the unit cell has two
opposite electric walls (zero electric field) and two magnetic walls (zero magnetic
field) with a TEM mode. However, when the beam is scanned, such a unit cell must
be tilted along with the beam, and another unit cell must be added for each grating
lobe. Because of this complexity Oliner utilised a fixed normal unit cell that would
accommodate any scan angle and any number of grating lobes. The unit cell walls
are not in general electric or magnetic, and the modes inside are LSE and/or LSM,
depending upon the scan plane. Opposite walls support fields that differ by the
Floquet phase shift. Each main lobe or grating lobe is represented by a propagating
mode, with the evanescent modes contributing energy storage (susceptance) at the
array face.

Fig. 10.59 Slot element and unit cell

Although short slots or dipoles are simpler to analyse by the periodic-cell


method, the resonant slot is of most importance; so this case will be carried through
in some detail. Scanning in a principal plane is also simpler but arbitrary scan will
be considered here. For this case all four unit cell walls are phase dependent, with
both Ez and Hz field components. A superposition of the dominant LSE and
dominant LSM modes is necessary. The wavenumber in the z direction is:
Planar arrays 205

K = vk2-klo-klo = kcosBo (10.111)


where
k sin e
= o cos 0O, kyo = k sin 6O sin (10.112)
The slot is located in a metal plane at the end of a feed waveguide. The array is then
represented as a junction between two waveguides: the feed guide and the unit cell
(see Fig. 10.60). Feed waveguide quantities are subscripted i for incident. Since a
slot has electric field in only one direction, the two unit cell modes can be replaced
by a single TE mode along the x direction (note the unit cell axis is z). The active
conductance becomes:
Yo]Vo
(10.113)
\Vt\2

cell
feed
wavegi

F ig. 10.60 Unit cell and feed guide

where the subscript o is for the dominant unit cell mode. Assuming a cosine slot
field with slot dimensions as, bs and feed waveguide dimensions a, b, the mode
voltages become:

kxoas . ky yobs
cose -££-=• sine (10.114)

where the unit cell dimensions are Dx = dx/X and Dy = dyj\\ the uniform excit-
ation space factor is:
sin A
sine A = (10.115)
206 Planar arrays

and the cosine space factor is:

These symbols simplify the formulas while allowing the various factors to be inter-
preted physically by inspection.
The active conductance, normalised to the feed guide characteristic admittance
is then:

(10.117)
2
2DxDyX> cose *f
where the scanning factor is:
1 — sin2 0O cos2Y
Hoo = ' "" T ° (10-118)
COS0 (

This will later be seen to be the dominant form of a more general scanning factor
Hnm-
For //-plane scan (scan parallel to the slots) it reduces to cos0 o while for
Z?-plane scan it becomes 1/cos 6O. This is in accord with the transmitting current-
sheet concepts presented earlier. The factors involving kxoas and kyobs are distribu-
tion factors representing the effects of the field distribution along and across the
slot, respectively. And the remaining factors can be separated into a factor for a
single waveguide fed slot and a mutual coupling factor:

3X3 cose 2 — -
2a
J
active _ ^ u
G
single slot

If the array lattice spacing is X/2 each axis, the active conductance is 3/TT times that
of a single slot. This, of course, is valid only at broadside where the scan factor is
unity.
When the combination of scan angles and lattice spacings allows one or more
grating lobes to appear, the conductance will have an added term for each GL. A
grating lobe appearing due to //-plane scanning sets up a higher TE mode in the
unit cell, and the slot conductance changes smoothly at grating lobe onset. This is
because the characteristic admittance of the mode is zero at GL onset. In contrast
the TM mode required by jE'-plane scanning has a characteristic admittance that is
infinite at GL onset, with the result that the conductance and VSWR have an
Planar arrays 207

infinite discontinuity there. This behaviour is related to the isolated element


pattern: in the plane of the slot the H-plane power is zero, yielding a smooth
grating-lobe transition. Conversely, the £-plane power is significant, producing a
discontinuity at onset. Dipole arrays are the dual; E- and //-plane behaviour is
reversed. However, a dipole array over a ground plane has an element pattern that is
zero in both E- and //-planes, and the active resistance transition is smooth in both
planes (Oliner and Malech76).
The slot engenders both internal and external susceptance. Internal susceptance
is discussed in Section 9.3.2.; it must be added to the external susceptance, which
has been found in a similar manner by Stark88 for dipole arrays, although his
development is not couched in unit-cell terms. The result is:

y ?D [ cosc 2 -^smc 2 -^// n m


2 2
1

(10.121)
J

The sums include all positive and negative integers except the term for n = m = o.
The generalised wave numbers are:

r n
kxn = A: sin 6O cos $ o 4- —
Dx
L
kym = k sin<9osin0o + ^ - (10.122)
[ Vy
and the generalized scan factor is:

»f
sin y o cos 0 O 4- —X/ I — 1
/ -2 ° 2 (10.123)
/(sin 0O cos 0 O H 1 4- j sin 0O sin <j)o H 1— 1
"V \ Ac/ \ ^y/
Instead of displaying active admittance data for various scan planes and spacings,
the active impedance of a dipole array will be discussed. The active slot admittance
and active dipole impedance will be seen to be simply related. Since most of the
available data have been computed for dipole arrays, it will be shown later. A dipole
array over a ground plane will be discussed as this includes the special case of no
ground plane.

10.3.3.2 Dipoles: The slot equivalent circuit is simple because the susceptance
is in parallel with both the feed waveguide and the unit-cell waveguide. Assuming
dipole feed terminals at the centre of the dipole, the reactance is in parallel with the
unit-cell waveguide, but in series with the feed terminals. Consideration of the
ground plane displays one of the advantages of the unit-cell approach. Whereas the
dipole array in free space is represented by a unit cell on each side of the dipole
208 Planar arrays

with both unit cells semi-infinite in length, the ground plane is simply represented
by a short across one of the unit cells at height h from the dipole. Since the ground
plane to dipole spacing is usually of the order of A/4, the higher-order unit-cell
evanescent modes are sufficiently damped out in travelling from the dipole to the
ground plane and back that they can be neglected. However, the ground plane
affects both resistance and reactance through the dominant mode. The dipoles are
assumed to be strip (flat) dipoles, so that the dipole array is the Babinet equivalent
of the slot array previously discussed. Fig. 10.61 shows the dimensions. The dipole
current distribution is assumed to flow only in the x-direction and to be constant
over the dipole in the y-direction. Along x a cosine distribution is assumed. A
match is then made between the electric field in the dipole plane produced by the
dipole, and that produced by a set of unit cell waveguide modes. From this, the
series for active reactance is obtained. For the active resistance the ground plane
is replaced by an image dipole at distance 2h. The results obtained originally by
Stark87 have been cast into the unit-cell format and symbology. Wave numbers
are the same as for the slot case. The active resistance is:
480a 2 kxoas kyobs
Rn = cose2 sine2 (sin2 kzoh)Ho (10.124)

J L

Dy = dy/X

Fig. 10.61 Flat dipole array dimensions

When a ground plane is not present, the sin2 kzoh factor is replaced by \. The
expression for reactance is:
240a2 2 JE £ 2 2
2 I cose - r- sine - ^ (sin kzoh)Ho
D
7tDxyJJyvA I I I

(10.125)
Planar arrays 209

Symbols are the same as before. Absence of the ground plane is handled by putting
h = 0.
Next, dipole active impedance is examined. Keep in mind that slot active admit-
tance can be obtained from the active impedance of a dipole without ground plane
simply by multiplying by a scale factor. Most of the data which follow were calcu-
lated by Diamond and are shown by courtesy of Oliner and Malech.76 Figs. 10.62
and 10.63 show active resistance normalised to unity at broadside for half-wave

3 r

E-piane

20 40 60 80
8 , deg

Fig. 10.62 Dipole-array active resistance

H-plane

0 20 40 60 80
8,deg

Fig. 10.63 Dipole array with ground-screen active resistance


D = 0.5X

dipoles in £-plane, //-plane and diagonal-plane scan. As mentioned, the //-plane


resistance goes to infinity as scan approaches 90°; for ^-plane scan the resistance
goes to zero for no ground plane. With the ground plane separated by a nominal
quarter wavelength from the dipoles, the resistance goes to zero for all scan planes.
210 Planar arrays

These two Figures are for half wave spacing, so no grating lobes occur. The next
two Figures, Figs. 10.64 and 10.65 show active resistance for dipoles over a ground
plane for lattice spacing of 0.6 and 0.7 wavelength. The solid curve is from periodic-
structure calculation, while the circles use the Carter element-by-element approach
on a 65 x 149 element array. The abrupt discontinuity in//-plane active resistance
is noteworthy; in the 2^-plane only modest change of direction occurs, while in the

220 -

200 -

180 -
1
I \ \
160 - / , H Plane \

140

-e-120

~ 100 /D Plane

- v
a
Q:
80

E Plane / \ \
60 Scant 0-0') \

40
Grating Lobe
Boundary—**
20
\
n i i i i

20 40 60 80
8,deg

Fig. 10.64 Dipole array over screen active resistance


o Carter
periodic structure
Dx = Dy = 0.6

diagonal plane the values are little affected. Agreement between the element-by-
element and periodic-structure approaches is excellent except for angles beyond
85° where the edge effects become significant. Since the dipole element pattern in
the //-plane does not go to zero, the edge effects are expected to be larger than for
/^-plane scan. Since curves for dipoles without ground screen, with element spacings
above 0.5X, have not been presented it should be recalled that they can be obtained
from the curves of dipoles with a ground screen by dividing out the 2 sin2 (kh cos 6)
factor. Note that the behaviour of active impedance with scan plane is a function of
Planar arrays 211
element pattern rather than element polarisation. Thus the singular behaviour is for
scanning normal to the radiating elements, which is //-plane for dipoles but is
/T-plane for slots.
Turning now to the active reactance, the computation involves a double infinite
series, and one must always be concerned with convergence questions. The periodic-
structure series converges slowly. When no ground plane is present, the element-by-
element calculation converges very slowly (a large array is required to approximate
the periodic structure results), but with a ground plane it converges more quickly.

180

160

UO

120

: 100

3
80'

60
D-plane scan
W

20

0,
20 40 60 80
e.deg
Fig. 10.65 Dipole array over screen active resistance
o Carter
periodic structure
Dx = Dy = 0.7

If the element-by-element calculation is used it should be done for the array on a


ground plane. The ground-plane effect can then be subtracted using the periodic-
structure term. Or the periodic-structure series can be used directly as the many
terms required are no longer serious with the high-speed computers of today. Figs.
10.66-10.68 show active reactance for dipoles in free space for half wavelength
spacing, for//-plane, diagonal-plane, and/T-plane scans. Curves are given for a dipole
width of 0.005 and 0.05 wavelength, and it can be seen that there is a modest
change at broadside in the //-plane, which decreases with scan and with rotation of
the scan plane. However, the current distribution across the dipole has been
assumed constant so fat dipoles should not be analysed by this method. The next
Figure Fig. 10.69, is for the same array (of thin dipoles) over a ground plane, where
active reactance is calculated two ways. First, the element-by-element method is
used with a sufficiently large array for convergence. Then the periodic-structure
series is used with various numbers of terms. Use of 441 terms (n = ± 21, m = ± 21)
gives results very close to the first values. Note that the ground plane does not con-
212 Planar arrays

trol the //-plane reactance singularity at grating-lobe incidence (90° for this case).
Figs. 10.70 and 10.71 show active reactance for lattice spacings of 0.6 and 0.7
wavelength; the singularity at GL onset is apparent.
From all these data it is apparent that arrays with large scan ranges will experi-
ence large changes in active-element impedance. Methods of reducing these changes
will be discussed later.

vO-005

005

-10

-20

Fig. 10.66 Dipole array reactance, H-piane scan showing dipole-width effects

250

200

150

100

50

0'
20 40 60 80
8,deg
Fig. 10.67 Dipole array reactance; E-plane scan showing dipole width effects

103.3.3 Blindness: A blind spot occurs in some arrays;it is actually an angle at


which the array radiates no power. This phenomenon can, like array performance in
general, be viewed from the standpoint of either active reflection coefficient or
active element pattern. The active reflection coefficient may develop a magnitude
near or equal to unity, or the active element pattern may be near or equal to zero.
Taking first the element active-reflective-coefficient viewpoint, the blind spot
Planar arrays 213

30

c j 20
X
10

80
0
20 40
0 , deg
-10

Fig. 10.68 Dipole array reactance; diagonal scan showing dipole width effects

180 /
/ ' '

160
III 1

140
!1 h11
H Plane Scan
1 Term
1« 1 9 Te r m s
III 1 101 TV} r rn T
N ' 441 To rm r

ill
a
j
100 a
X
o ill1 '
in i
80 n i
11 i
I i
60
/" '
Jr / /
E Plane Scan
1 Torm
40 '^y' / / / / — 121 Terms
— 441 Terms

0
20 40 60 80 100
e deg

Fig. 10.69 Dipole array with ground screen active reactance, showing effect of number of
terms
element-by-element method
periodic structure method
. = D v = 0.5
214 Planar arrays

200

180

160

H-plane scan
U0 grating lobe
(0=90°) boundary
120 D-plane
scan

Fig. 10.70 Dipole array over screen active reactance


o Carter
periodic structure
Dx = Dy = 0.6

180

160

D plane scan(<J>=45°)
UO

120

t
x 80
/ H plane scan
(4>=90°)
60 /
40 E plane scan (c^zO^ 0 ^

20* -
^ g r a t i n g lobe appears
1 1 i I i I i i
°c) 20 40 50 80
0. deg

Fig. 10.71 Dipole array over screen active reactance


o Carter
periodic structure
Dx = Dy = 0.7
Planar arrays 215
occurs when a higher mode cancels the dominant mode in the element. Since the
relative phase changes rapidly with scan angle, the two modes produce a resonance,
with a resonant <p that can be defined on the decrease of | F | with scan angle. The
higher mode can be produced by external or internal structure. Examples of
external structure causing a blind spot are a dielectric radome covering each
element or row of elements (Lechtreck63), a dielectric sheet placed on the array
face (Bates;8 Byron and Frank;13 Wu and Galindo;110 Gregorwich, Hessel, Knittel
and Oliner34), and a dielectric plug protruding from the mouth of each element
(waveguide) (Hannan41). In all cases the active wave impedance at the array face is
modified by the external structure so as to produce a dominant-mode, higher-mode
resonance at a particular scan angle; the higher mode exists primarily outside the
face. Examples of internal structure that can allow a higher mode that can cause a
blind spot are dielectric plugs flush with the array face (Amitay and Galindo4),
die lee trie-loaded waveguide elements (as opposed to a short plug) (Diamond20), and
a brick array of rectangular guide elements (Diamond20). The latter is sketched in
Fig. 10.72. For these various structures the waveguide element supports the next
higher mode, which causes the resonance at the blind angle. For the brick array the
higher mode is just below cutoff due to the large guide size. An equivalent circuit
for the two modes has been used by Oliner74 and Knittel et al.5S to predict the
reflection coefficient behaviour. Complete reflection corresponds to an active con-
ductance that is zero. Simulators (see Section 10.3.3.4.) have been used both to
verify the presence of blind spots, and to design them out.

Fig. 10.72 Brick waveguide array

From the active-element-pattern viewpoint, the active impedance at the array


face, including the effect of external or internal structure, allows a leaky wave to
be excited. This leaky wave is usually backward, and its amplitude depends on the
active impedance and on the length of the impedance surface (the array face). Thus
small arrays usually have no blind spot or only a small dip instead of a null, while
very large arrays are more susceptible as it is easier to excite the leaky wave. The
leaky wave has a complex wavenumber, analogous to Wood anomalies on an
optical-reflection grating (Hessel and Oliner48). When the angle and phase of leaky
waves are suitable, the direct radiation at that angle will be cancelled, providing an
216 Planar arrays

active-element pattern null, or blind spot. The leaky-wave pole gives an approxi-
mate behaviour of cancellation around the blind angle, where F is expanded into a
Taylor series about the pole. Fig. 10.73 shows the active reflection coefficient for
an array of narrow slots fed by parallel-plate waveguides and covered by a dielectric
slab, for iT-plane scan. This curve was calculated using 21 modes. Slot spacing was
X/2, waveguide width was 0.3X, while the slab had e = 2.56 and thickness 0.2X.
Using the leaky-wave pole Knittel et al.59 calculated the dots shown in the Figure,
where the leaky-wave Q is given by:
u2 Re (kjk)
2 (10.126)
l-u 2lm(kx/k)

i-o

from leaky wave

0-8

0-6

0-4

0-2

0 0-2 0-4 0-6 0-8 10


sin 6

Fig. 10.73 Blind spot of array of slits (courtesy: Knittel, Hessel and Oliner)

kx is the wavenumber along the array. For this array the grating lobe occurs at
90°, and the active reflection coefficient is also unity there. Another example
used a brick array of rectangular guides 0.905 by 0.4X, on a square brick lattice
of 1.008X spacing. An array of only 7 x 7 elements exhibits a blind spot over
26 dB deep in the //-plane at 27° (Farrell and Kuhn26 ). A grating lobe occurs on
the 60° circle in u-v space; the blind spot is usually just inside the grating lobe,
but in this example it is well inside (Oliner75). Other discussions of blind spots are
in Section 10.3.2.
The understanding of how blind spots arise also gives indications of corrective
measures to take (Knittel;58 Oliner75). These include reducing dimensions to
Planar arrays 217
preclude/eliminate higher propagating modes, alternating external structure or
external impedance, and changing various parameters to obtain a suitably low
VSWR over a wide scan-angle range.

103.3.4 Simulators: The development of waveguide simulators allows theoreti-


cal designs to be verified at low cost. Such a simulator in simplest form consists of a
metal ground plane containing at least one and part of another slot, with an open-end
waveguide connected to one side of the ground plane (see Fig. 10.74). The wave-
guide dimensions are chosen so that the walls lie in lattice planes of symmetry.
The example shows a triangular and a rectangular lattice. Then the frequency is
selected so that the element spacing is correct and so the virtual plane-wave* com-
ponents of the waveguide mode make an angle with the guide axis that equals the
desired scan angle (Hannan and Balfour;38 Balfour;7 Wheeler105). Clearly such a
simulator is a 'one point' device in that the element spacing is correct at only one
frequency, and the waveguide (lattice) size then determines the scan angle. For slots
the waveguide can be fed directly. For open-end waveguide elements the main
element (hole) can be fed by an actual feed waveguide, complete with dielectric
plugs, matching height or width changes, etc. Partial elements are then terminated
with a short-circuit waveguide section; it is only necessary that they support reactive
modes. The feed waveguide excites the element, while the other elements provide
the proper reactive environment through symmetry to simulate active admittance,
which is then measured in the feed guide. Such simulators are only practical in the
absence of grating lobes.

Fig. 10.74 Wa veguide array Simula tors


a Triangular lattice b Rectangular lattice
Simplest useful simulators — d « 40° /V-plane

A more general purpose simulator was developed by Gustincic35 and Derneryd


and Gustincic,19 which allows active admittance at various scan angles to be deter-
mined. Grating lobes can also be encompassed. The simulator uses many elements,
with all elements and partial elements but two partially terminated. Transmission
measurements are then made between these two elements. In turn, the two active
elements occupy all positions in the simulator, which allows active admittance data
to be calculated for TV scan angles, if there are N simulator elements.

* The dominant TE mode in a rectangular waveguide can be decomposed into two plane waves
that reflect alternately from the narrow walls
218 Planar arrays
103 A Scan compensation

10.3AJ Multi-mode elements: From the earlier discussion on ideal element


patterns it can be inferred that both TE and TM modes are necessary in the
periodic-structure unit cell to produce a reflection coefficient that varies less with
scan. Probably the most effective way of producing the optimum mix of modes
uses open-end circular or rectangular waveguide elements, with a dielectric plug
inside each guide at the face, and a dielectric sheet over the array face. The periodic-
structure unit cell can be used to design such wide-angle impedance-matched
elements, as described in Section 9.3.3. There are variations on this technique,
where each element has an iris and a dielectric plug at its mouth to excite higher
modes (Wong et al.10S), where a dielectric sheet is placed on the array face to
improve //-plane scan (by broadening the element pattern) (Galindo and Wu29), or
where a thin high-dielectric-constant sheet is placed near the array face to cancel
some array-scan susceptance (Magill and Wheeler67). A survey of wide-angle
impedance matching is given by Knittel.58
Schemes for reducing mutual coupling rather than compensating the variation
with scan angle are also useful. The simplest way is to pack the elements closely
together in wavelengths. The results in the previous Section showed that the admit-
tance variation with scan increases as grating-lobe onset is approached. Reducing
the element spacing below half wave reduces the admittance variation owing to the
GL being farther away (in invisible space). This is, of course, difficult without going
to narrow-band elements. Another scheme changes the impedance of the ground
plane between elements by introducing corrugations (Dufort22). There is a tradeoff
here in that it may be necessary to increase element spacing to allow room for
corrugations. A related but earlier scheme uses //-plane baffles between slots or
dipoles (Oliner and Malech77). These baffles are simply metal strips perpendicular
to the ground plane and running between the dipole rows in the //-plane
(Mailloux68). Fig. 10.75 from Edelberg and Oliner24 shows the improvement that
can be realised. In this example the dipoles are spaced 0.5X apart and 0.25X above a
ground plane. Use of baffles about 0.5 X high makes the E- and//-plane patterns
closely alike. Over a 60° conical scan volume a VSWR < 1.6 can be achieved.
Another scheme modifies the element patterns directly. Fig. 10.76 shows a
longitudinal shunt slot array with parasitic dipoles developed by Clavin et al.ls The
dipoles reduce the /T-plane beamwidth and add an electric-type mode to comple-
ment the slot magnetic-type mode. Patterns in E- and//-planes can be made equal.
Mutual coupling between elements is reduced 14 dB, and the array wide-angle side-
lobes and backlobe are significantly reduced.
Network methods of impedance matching are discussed in the Section on
element-by-element approaches.

103A.2 Active element pattern modified by dielectric sheets: It has been


repeatedly observed in this Chapter that the ideal element pattern is rotationally
symmetric, and is cos 6 over most if not all of the polar angle 9 from 0 to TT/2.
Planar arrays 219
Since most array elements have E- and //-plane patterns that are unequal, effecting
a wide-angle impedance match requires altering one or both of these patterns.
Excitation of multiple modes to approximate a cos 6 pattern has been discussed.

H plane
(unaffected)

0 15 30 45 60 75 90 105 120 135


scan angle, deg

Fig. 10.75 Active resistance of dipole array over ground plane with baffles {courtesy:
Edelberg and Oliner)
without baffles
— with baffles

Fig. 10.76 Slot with parasitic elements {courtesy: Clavin, Huebner and Kilburg)

It is also possible to effect the required changes through use of an external spatial
filter (Mailloux69). Here, of course, the spatial filter not only has a transfer function
220 Planar arrays

that varies with 0, but also with polarisation. Such filters have been developed by
Munk and colleagues.71'72 The filter consists of several contiguous flat dielectric
sheets, placed over the array face. Since the plane-wave transmission coefficient of
a dielectric sheet varies with both angle and polarisation the potential is obvious.
To analyse this configuration, and to design a compensation unit, the field from an
infinite 2-dimensional array of elements is expanded into a 2-dimensional spectrum
of plane waves, weighted by the open-circuited element pattern. A single plane-wave
component has a transmission coefficient through the sheets that is obtained from
the cascading of impedances. Each sheet has its own thickness and dielectric con-
stant, and low loss; the transmission coefficient is r, found from the impedance
cascade. The sum of the weighted plane waves over the array elements times the
transmission coefficient gives the radiated field (active element pattern); if the
sheet parameters are optimally chosen the result might approach cos 6. The field
of a uniformly excited array of dipoles, for example, is:
JZ

Here the dipoles are along x in the x-y plane; h is the isolated dipole pattern; for
details of the wavenumbers see Section 10.3.3. The scan vector i$axu + ayv, where
the aX9 ay are unit vectors. This bundle of inhomogeneous plane waves is propagat-
ing when kz is real; for imaginary kz the waves are evanescent. / is the realised
dipole current.
The transmission coefficient r of an exfoliated dielectric as mentioned is found
from the cascaded impedances; a simple way of calculating this cascade uses an
A-B-C-D matrix or T matrix for each layer, since the matrix of the cascade is the
product of the matrices of each layer. The T matrix is loosely related to trans-
mission coefficients, so it is used here. With slabs 1-rc, with slab 1 at the array face,
the overall T matrix is:

T = 7V T V , . . . TV TV (10.128)
However, more than one T matrix is required per slab, so the impedance cascade
contains more factors. The interface from free space to a slab, or from one slab to
a contiguous slab, requires an impedance transformation matrix. A transmission
line also requires a matrix. Thus n slabs utilise a cascade with n line matrices inter-
spersed among n 4- 1 transformation matrices. When the transmission line is a
dielectric slab, the incidence angle and polarisation must be included. The con-
stituent matrices are (Montgomery eta/.;70 Ramo et a/. 79 ):

transformation matrix T = (10.129)

-i(H KH.
Planar arrays 221

where the transformation ratio n is:

e cos 6
r = parallel polarisation (10.130)
— sin 2

COS0
= , ==r normal polarisation
Ve-sin 2 0
lossless
dielectric T = (10.131)
slab

Obviously the slab thickness is / and the angle of propagation from normal is $.
Given the parameters of the sheets, the overall T matrix is easily computed. The
transmission coefficient r = S21 a n d f° r reciprocal slabs S2i = Si2« Since T22 =
I/S12 the transmission coefficient is immediately:

r = (10.132)

Although only one element of T is needed, each constituent T matrix must be


carried in complete form until the final product is obtained.
As an example for a single slab, the two transformation and one slab matrices
are multiplied, from which T22 is easily found. The transmission coefficient reduces
to:

T = (10.133)
1 - r2 exp (-/2/c/Ve-sin 2 0)

Here r is the air-dielectric interface Fresnel reflection coefficient. Note that the
insertion-phase delay term which allows a virtual straight-line path is not included
inr.
Returning to the electric-field expression, r is inserted into the double sum.
However, for general scan, the polarisation of the elements must be resolved into
parallel and normal components, with each multiplied by an appropriate r and
element pattern. If now the dielectric slabs can be chosen to make the sum approxi-
mately cos# over a scan range, the array is compensated. Munk, Kornbau and
Fulton have matched an array of short waveguide slots by a single dielectric slab of
e = 1.33 and r/X 0 = 0.367, with a VSWR < 1.5 for 6 between 0 and 80° in both
principal planes. This assumes that the self admittance of the slot has been matched
separately at broadside; the slab produces a match vs. scan angle. Work is in progress
on slab matching of resonant slots and dipoles. At the stage of writing, the band-
width behaviour has not been determined. This technique is significant and promis-
ing in that it utilises single-mode elements, which are simple and inexpensive.
222 Planar arrays
10.4 References

1 ABRAMOWITZ, M., and STEGUN, I. A.: 'Handbook of mathematical functions'. NBS,


AMS-55,1970
2 ALLEN, J. L.: 'Gain and impedance variation in scanned dipole arrays', IRE Trans.,
AP-10, 1962, pp. 566-572
3 AMITAY, N., et al.: 'On mutual coupling and matching conditions in large planar phased
arrays', IEEE G-AP Int. Symposium Digest, 1964, pp. 150-156
4 AMITAY, N., and GALINDO, V.: 'Analysis of circular waveguide phased arrays', BSTJ,
47, 1968, pp. 1903-1931
5 ANDERSEN, J. B., et al.: 'Coupling between minimum scattering antennas', IEEE Trans.,
AP-22, 1974, pp. 832-835
6 ANDERSEN, J. B., and RASMUSSEN, H. H.: 'Decoupling and descattering networks for
antennas',IEEE Trans., AP-24, 1976, pp. 841-846
7 BALFOUR, M. A.: 'Active impedance of a phased-array antenna element simulated by a
single element in waveguide', IEEE Trans., AP-15, 1967, pp. 313-314
8 BATES, R. H. T.: 'Mode theory approach to arrays', IEEE Trans., AP-13, 1965, pp. 321-
322
9 BEGOVICH, N. A.: 'Slot radiators',Proc. IRE, 38, 1950, pp. 803-806
10 BOOKER, H. G.: 'Slot aerials and their relation to complementary wire aerials (Babinet's
principle)', /. IEE, 93, Pt. IIIA, 1946, pp. 620-626
11 BORGIOTTI, G. V.: 'Radiation and reactive energy of aperture antennas', IEEE Trans.,
AP-11, 1963, pp. 94-95
12 BORGIOTTI, G. V.: 'A novel expression for the mutual admittance of planar radiating
elements', IEEE Trans., AP-16, 1968, pp. 329-333
13 BYRON, E. V., and FRANK, J.: 'Lost beams' from a dielectric covered phased-array aper-
ture', IEEE Trans., AP-16, 1968, pp. 496-499
14 CARTER, P. S.: 'Circuit relations in radiating systems and applications to antenna prob-
lems', Proc. IRE, 20, 1932, pp. 1004-1041
15 CLAVIN, A., et al.: 'An improved element for use in array antennas',IEEE Trans., AP-22,
1974,pp.521-526
16 COLLINS, J. R., and HAYES, W. F.: 'The RASSR array'. Proc. 1972 USAF Antenna
Symposium, Allerton, IL
17 CONTI, R., and TOTH, J.: 'The wire grid microstrip antenna', Proc. 1980 Antenna Appli-
cations Symp., Allerton, IL
18 DEBSKI, T. R., and HANNAN, P. W.: 'Complex mutual coupling measurements in a large
phased array antenna', Microwave J., June 1965, pp. 93-96
19 DERNERYD, A. G., and GUSTINCIC, J. J.: 'The interpolation of general active array
impedance from multielement simulators', IEEE Trans., AP-27, 1979, pp. 68-71
20 DIAMOND, B. L.: 'Resonance phenomena in waveguide arrays', IEEE G-AP Intl. Symp.
Digest, 1967, pp. 110-115
21 DIAMOND, B. L.: 'A generalized approach to the analysis of in finite planar array antennas',
Proc. IEEE, 56, 1968, pp. 1837-1850
22 DUFORT, E. C : 'A design procedure for matching volumetrically scanned waveguide
arrays', Proc. IEEE, 56, 1968, pp. 1851-1860
23 EDELBERG, S., and OLINER, A. A.: 'Mutual coupling effects in large antenna arrays.
Part I: Slot arrays', IRE Trans., AP-8, 1960, pp. 286-297
24 EDELBERG, S., and OLINER, A. A.: 'Mutual coupling effects in large antenna arrays.
Part II: Compensation effects', IRE Trans., AP-8, 1960, pp. 360-367
25 ELLIOTT, R. S.: 'The theory of antenna arrays', in HANSEN, R. C. (Ed.): 'Microwave
scanning antennas. Vol. IF., Academic Press, 1966, Chap. 1
26 FARRELL, Jr., G. F., and KUHN, D. H.: 'Mutual coupling effects in infinite planar arrays
of rectangular waveguide horns', IEEE G-AP Int. Symp. Dig., 1966, pp. 392-397
Planar arravs
27 GALINDO, V., and WU, C. P.: 'Numerical solutions for an infinite phased array of rect-
angular waveguides with thick walls',IEEE Trans., AP-14, 1966, pp. 149-158
28 GALINDO, V., and WU, C. P.; 'Dielectric loaded and covered rectangular waveguide phased
arrays', BSTJ, 47, 1968, pp. 93-116
29 GALINDO, V., and WU, C. P.: 'Surface wave effects on phased arrays of rectangular wave-
guides loaded with dielectric plugs', IEEE Trans., AP-16, 1968, pp. 358-360
30 GALLEGRO, A. D.: 'Mutual coupling and edge effects in linear phased arrays', MS Thesis,
Polytech. Inst. of Brooklyn, 1969
31 GATELY, Jr., A. C , et al.: 'A network description for antenna problems', Proc. IEEE, 56,
1968,pp.1181-1193
32 GOEBELS, F. J., and KELLY, K. C : 'Arbitrary polarization from annular slot planar
antennas',//?£ Trans., AP-9, 1961, pp. 342-349
33 GOLDBERG, J. J.: 'Study of a phased array finite in its plane of scan'. MS Thesis, Poly-
tech. Inst. of Brooklyn, 1972
34 GREGORWICH, W. S., et al.: 'A waveguide simulator for the determination of a phased-
array resonance'. IEEE G-AP Intl. Symp. Digest, 1968, pp. 134-141
35 GUSTINCIC, J. J.: The determination of active array impedance with multielement wave-
guide simulators', IEEE Trans., AP-20, 1972, pp. 589-595
36 HAMID, M. A. K.: 'Mutual coupling between sectoral horns side by side', IEEE Trans.,
AP-15, 1967, pp. 475-477
37 H ANN AN, P. W.: 'The element-gain paradox for a phased array antenna', IEEE Trans.,
AP42, 1964, pp. 423-433
38 HANNAN, P. W., and BALFOUR, M. A.: 'Simulation of a phased-array antenna in wave-
guide\ IEEE Trans., AV-13, 1965, pp. 342-353
39 HANNAN, P. W.: 'The ultimate decay of mutual coupling in a planar array antenna',
IEEE Trans., AP-14, 1966, pp. 246-248
40 HANNAN, P. W;.: 'Proof that a phased-array antenna can be impedance matched for all
scan angles', Radio Sci., 2, 1967, pp. 361-369
41 HANNAN, P. W.: 'Discovery of an array surface wave in a simulator',IEEE Trans., AP-15,
1967,pp.574-576
42 HANSEN, R. C : 'Tables of Taylor distributions for circular aperture antennas',IRE Trans.,
AP-8, 1960, pp. 23-26
43 HANSEN, R. C : 'Microwave scanning antennas; Vol. 1', Academic Press, 1964, Chap. 1.
44 HANSEN, R. C : 'Comparison of square array directivity formulas', IEEE Trans., AP-20,
1972,pp. 100-102
45 HANSEN, R. C : 'Formulation of echelon dipole mutual impedance for computer', IEEE
Trans., AP-20, 1972, pp. 780-781
46 HANSEN, R. C : 'A one-parameter circular aperture distribution with narrow beamwidth
and low sidelobes', IEEE Trans., AP-24, 1976, pp. 477-480
47 HANSEN, R. C , and BRUNER, G.: 'Dipole mutual impedance for design of slot arrays',
Microwave J., 22, Dec. 1979, pp. 54-56
48 HESSEL, A., and OLINER, A. A.: 'A new theory of Wood's anomalies on optical gratings',
Appl. Optics, 4, 1965, pp. 1275-1297
49 IMMELL, R. G., and SASSER, B. If: 'A highly thinned array using the image element
antenna', Proc. 1979 Antenna Appl. Symp., Allerton, IL.
50 INCE, W. J., and TEMME, D. H.: Thasers and time delay elements', in YOUNG, L. (Ed.):
'Advances in microwaves: Vol. 4'. Academic Press, 1969.
51 JEDLICKA, R. P., and CARVER, K. R.: 'Mutual coupling between microstrip antennas'.
Printed Circuit Antenna Technology Workshop, New Mexico State Univ., 1979, pp.
4-1 to 4-9
52 KAHN, W. K.: 'Ideal efficiency of a radiating element in an infinite array', IEEE Trans.,
AP-15, 1967, pp. 534-538
53 KAHN, W. K.: impedance-match and element-pattern constraints for finite arrays',IEEE
224 Planar arrays

Trans., AP-25, 1977, pp. 747-755


54 KAHN, W. K., and KURSS, H.: 'Minimum-scattering antennas', IEEE Trans., AP-13, 1965,
pp.671-675
55 KELLY, K. C, and GOEBELS, F. J.: 'Annular slot monopulse antenna arrays*, IEEE
Trans., AP-12, 1964, pp. 391-403
56 KING, H. E.: 'Mutual impedance of unequal length antennas in echelon', IRE Trans., AP-5,
1957,pp.300-313
57 KING, M. J., and THOMAS, R. K.: 'Gain of large scanned arrays', IRE Trans., AP-8, 1960,
pp.635-636
58 KNITTEL, G. H.: 'Wide-angle impedance matching of phased-array antennas: A survey of
theory and practice', in OLINER, A. A., and KNITTEL, G. H. (Eds.): 'Phased array
antennas'. Artech House, 1972
59 KNITTEL, G. H., et al.\ 'Element pattern nulls in phased arrays and their relation to
guided waves', Proc. IEEE, 56, 1968, pp. 1822-1836
60 KRAUS, J. D.: 'Antennas'. McGraw-Hill, 1950
61 KRAUS, J. D.: 'A backward angle end fire antenna', IEEE Trans., AP-12, 1964, pp. 58-60
62 KROWNE, C. M., and SINDORIS, A. R.: '#-plane coupling between rectangular microstrip
antennas', Electron. Lett., 16, 1980, pp. 211-213
63 LECHTRECK, I. W.: 'Effect of coupling accumulation in antenna arrays', IEEE Trans.,
AP-16, 1968, pp. 31-37
64 LO, Y. T., and LEE, S. W.: 'Affine transformation and its application to antenna arrays',
IEEE Trans., AP-13, 1965, pp. 890-896
65 LO, Y. T.: 'A note on the cylindrical antenna of noncircular cross section', /. Appt.
Phys., 24,1953, pp. 1338-1339
66 LYON, J. A. M., et ah: 'Interference coupling factors for pairs of antennas', Proc. 1964
USAF Antenna Symposium, Allterton, IL.
67 MAGILL, E. G., and WHEELER, H. A.: 'Wide-angle impedance matching of a planar
array antenna by a dielectric sheet', IEEE Trans., AP-14, 1966, pp. 49-53
68 MAILLOUX, R. J.: 'Reduction of mutual coupling using perfectly conducting fences',
IEEE Trans., AP-19, 1971, pp. 166-173
69 MAILLOUX, R. J.: 'Synthesis of spatial filters with Chebyshev characteristics', IEEE
Trans., AV-24, 1976, pp. 174-187
70 MONTGOMERY, C. G. (Ed.): 'Technique of microwave measurements'. McGraw-Hill,
1947
71 MUNK, B. A., and BURRELL, G. A.: 'Plane-wave expansion for arrays of arbitrarily
oriented piecewise linear elements and its application in determining the impedance
of a single linear antenna in a lossy half-space', IEEE Trans., AP-27, 1979, pp. 331-
343
72 MUNK, B. A., et ah: 'Scan independent phased arrays', Radio Sci.3 14, 1979, pp. 979-
990
73 NELSON, E. A.: 'Quantization sidelobes of a phased array with a triangular element
arrangement', IEEE Trans., AP-17, 1969, pp. 363-365
74 OLINER, A. A.: 'On blindness in large phased arrays'. URSI Symp. on Electromagnetic
Waves, Stresa, Italy. Alta Freq., 38,1969, pp. 221-228
75 OLINER, A. A.: 'Surface-wave effects and blindness in phased-array antennas', in
OLINER, A. A., and KNITTEL, G. H.: 'Phased array antennas', Artech House, 1972
76 OLINER, A. A., and MALECH, R. G.: 'Mutual coupling in infinite scanning arrays', in
HANSEN, R. C. (Ed.): 'Microwave scanning antennas: Vol. II'. Academic Press, 1966,
Chap. 3
77 OLINER., A. A., and MALECH, R. G.: 'Mutual coupling in finite scanning arrays', in
HANSEN, R. C. (Ed.): 'Microwave Scanning Antennas: Vol. II'. Academic Press,
1966, Chap. 4
78 OPP, F. L.} et aL: 'RASSR analysis, test and evaluation program'. Technical Report
Planar arrays 225

AFAL-TR-7744, Texas Instr., July 1977


79 RAMO, S., et al.\ 'Fields and waves in communication electronics'. John Wiley, 1965
80 REED, J. E.: 'The AN/FPS-85 radar system', Proc. IEEE, 56, 1969, pp. 324-335
81 RHODES, D. R.: *On a fundamental principle in the theory of planar antennas', Proc.
IEEE, 52,1965, pp. 1013-1021
82 RHODES, D. R.: 'Synthesis of planar antenna sources'. Clarendon Press, 1974
83 RICHMOND, J. H.: 'Mutual impedance between coplanar-skew dipoles', IEEE Trans.,
AP-18, 1970, pp. 414-416
84 RUDDUCK, R. C, et al.: 'Directive gain of circular Taylor patterns', Radio Sci., 6, 1971,
pp.1117-1121
85 SCHELKUNOFF, S. A., and FRHS, H. T.: 'Antenna theory and practice', Wiley, NY,
1942, pp. 368 and 401
86 SHARP, E. D.: 'A triangular arrangement of planar-array elements that reduces the
number needed', IRE Trans., AP-9, 1961, pp. 126-129
87 STARK, L.: 'Radiation impedance of a dipole in infinite arrays', Hughes Aircraft Co.,
Tech. Rept. FL60-230,1960
88 STARK, L.: 'Radiation impedance of a dipole in an infinite planar phased array', Radio
Sci.A, 1966, pp. 361-377
89 STEYSKAL, H.: 'Mutual coupling analysis of a finite planar waveguide array',IEEE Trans.,
AP-22, 1974, pp. 594-597
90 TAYLOR, T. T.: 'Design of circular apertures for narrow beam widths and low sidelobes',
IRE Trans., AP-8,1960, pp. 17-22
91 TIURI, U.,etaL: 'Chain antenna', 1974 IEEE APS Symp. Digest, pp. 274-277
92 TSANDOULAS, G. N.: 'Tolerance control in an array antenna', Microwave /., 20, Oct.
1977, pp. 24-30.
93 VARON, D., and ZYSMAN, G. I.: 'On the mismatch of electronically, steerable phased-
array antennas', Radio Sci., 3, 1968, pp. 487-489
94 VONAULOCK, W. H.: 'Properties of phased arrays', Proc. IRE, 48, 1960, pp. 1715-
1727
95 VON TRENTINI, G.: 'Partially reflecting sheet arrays', IRE Trans., AP-4,1956, pp. 666-
671
96 WASYLKIWSKYJ, W., and KAHN, W. K.: 'Theory of mutual coupling among minimum-
scattering antennas', IEEE Trans., AP-18, 1970, pp. 204-216
97 WASYLKIWSKYJ, W., and KAHN, W. K.: 'Efficiency as a measure of size of a phased-
array antenna', IEEE Trans., AP-21, 1973, pp. 879-884
98 WASYLKIWSKYJ, W., and KAHN, W. K.: 'Element patterns and active reflection
coefficient in uniform phased arrays', IEEE Trans., AP-22, 1974, pp. 207-212
99 WASYLKIWSKYJ, W., and KAHN, W. K.: 'Element pattern bounds in uniform phased
arrays', IEEE Trans., AP-25, 1977, pp. 597-604
100 WESTLAKE, J. R.: 'A handbook of numerical matrix inversion and solution of linear
equations'. John Wiley, 1968, Chaps. 2 and 7
101 WHEELER, H. A.: 'The radiation resistance of an antenna in an infinite array or wave-
guide', Proc. IRE, 36,1948, pp. 478-488
102 WHEELER, H. A.: 'The radiansphere around a small antenna', Proc. IRE, AP-47, 1959,
pp.1325-1331
103 WHEELER, H. A.: 'Simple relations derived from a phased-array antenna made of an
infinite current sheet', IEEE Trans., AP-13, 1965, pp. 506-514
104 WHEELER, H. A.: 'The grating-lobe series for the impedance variation in a planar phased-
array antenna',IEEE Trans., AP-14,1966, pp. 707-714
105 WHEELER, H. A.: 'A survey of the simulator technique for designing a radiating element',
in OLINER, A. A., and KNITTEL, G. H. (Eds.): 'Phased array antennas', Artech
House, 1972, pp. 132-148
106 WIMP, J.: 'Polynomial approximations to integral transforms', Math. Tables Other Aids
226 Planar arrays
Compute 15, 1961, pp. 174-178
107 WOLFSON, R. L, and CHO, C. F.: 'A wideband, low-sidelobe, polarisation-agile array
antenna', IEEE Int. Radar Conf. Record, 1980, pp. 284-287
108 WONG, N. S., et al.: 'Multimode phased array element for wide scan angle impedance
matching1, in OLINER, A. A., and KNITTEL, G. H. (Eds.): 'Phased array antennas',
Artech House, 1972, pp. 178-186
109 WOODY, W. C , and HANSEN, R. C : Paper to be published
110 WU, C. P., and GALINDO, V.: 'Surface wave effects on dielectric sheathed phased arrays
of rectangular waveguides', BSTJ, 47,1968, pp. 117-142.
111 ZAGHLOUL, A. I., and MACPHIE, R. H.: 'Generalized cross-correlation analysis of
phased array admittance', Radio Sci., 11,1976, pp. 61-70
112 ZAGHLOUL, A. I., and MACPHIE, R. H.: 'Analysis of finite phased arrays of narrow slots
using correlation matrix method', IEEE Trans., AP-27,1979, pp. 261-264
Chapter 11

Conformal arrays
Giorgio V. Borgiotti

11.1 Introduction

Conformal arrays can be defined in the most general terms as arrays that are non-
planar. However, often a more restrictive definition is used that postulates an array
whose elements are flush mounted on a non-planar surface.12 It is implied in this
definition, adopted here, that the presence of the surface on which the elements are
located plays an important role, affecting the radiative and impedance properties of
the array. Accordingly, the study of non-planar radiating arrays in free space, such
as the circular arrays of dipoles, a subject of thorough investigations in the past,22
is not considered to be within the scope of this Chapter.
Conformal arrays are, in general, more difficult to analyse than their planar
counterparts. Notwithstanding, conformal arrays are considered attractive
alternatives for certain applications for which planar arrays have definite draw-
backs. Specifically, great potential exists in the following two areas:
(a) Ground-based systems, requiring a wide angle or omniazimuthal coverage
(b) Arrays mounted on missiles or high-speed aircraft.
The maximum scan angle from broadside for a planar array cannot be made much
greater than 60°, because of the gain losses due to both the reduction of the pro-
jected aperture in the scan direction and the aperture mismatch. Consequently,
obtaining a wide-angle coverage by using planar arrays requires the use of a multi-
face system. This solution has been adopted in ground radar systems (Fig. 11.1).
However, the potential offered by conformal arrays in this area is tremendous, a
conformal array on a spherical surface (Fig. 11.2) or on a wide aperture cone being
in principle a solution better matched to the geometry of the problem.
For missile borne and airborne applications, conformally flush-mounted arrays
have the advantage of not introducing any additional aerodynamic drag and also of
providing wide-angle coverage, if so required. Other advantages are:
• Less space is required in the missile body, because the radiating elements are
pushed far forward in the missile nose, leaving space to package the phase
shifters and control circuitry.
228 Conformal arrays

Potential increase of the available aperture — as there is no swept volume. This


can be used to provide narrower beam widths and possibly higher antenna gain.
Elimination or reduction of the radome-induced boresight error. For an antenna
in a radome, there is an interaction between the two which depends on the
antenna position and beam-steering condition. In a conformal array, either the
radome is eliminated completely or the elements are close to the radome, with a
substantial reduction of the pattern distortion..

Fig. 11.1 Planar array multiface system for wide-angle coverage (courtesy: Raytheon Co.)

There exists a number of difficult engineering problems in conformal array design.


Some of them are listed below, although their discussion is beyond the scope of this
chapter:
• Maintaining array polarisation. Because the conformal array is three-dimensional,
it is difficult to ensure that all the elements radiate the same polarisation in the
angular region of interest. Circularly-polarised elementS4)rovide a partial answer,
at least for certain applications. However, obtaining an element pattern circu-
larly polarised over a wide angle is not an easy task.
• The elements are located on singly or doubly curved surfaces. It is difficult to
deposit or etch elements on curved surfaces. For a singly curved surface, which
can be enrolled on a planar surface, e.g. a cone, the problem is, of course,
simpler.
Con formal arrnvs
Packaging and feed system. If the widest angular coverage compatible with the
conformal array geometry is sought, it is necessary to control the power distri-
bution via a system of microwave switches. This makes the design of the feed
system more difficult than for the planar case. Also, to make effective use of the
volume saving that a conformal array provides, the phase shifters will have
differing orientations, with resulting difficult interconnection problems. Con-
formal arrays could have severe topological problems in both RF and DC control
areas.

Fig. 11.2 Spherical dome array (courtesy: Sperry Gyroscope)

11.2 Conformal array analysis

11.2.1 General
The tremendous progress of microwave component technology in recent years,
particularly in the area of phase shifters, has made phased arrays a practical solution
for scanning antenna systems. For planar arrays effective analytical tools have been
developed for performance prediction, especially for the evaluation of the gain
230 Con forma/ arrays

vis-a-vis scan angle. On the other hand, methods of analysis and synthesis of con-
formal arrays have not been developed to the same extent as their counterparts for
planar structures.
The analysis of the electromagnetic properties of a conformal array is a problem
whose mathematical structure consists of two essentially separate parts. One is the
classical 'external' electromagnetic problem, that is the determination of the
electromagnetic field in the region exterior to a closed surface on which known
electric or magnetic currents exist. The solution of this problem is known for
simple geometries. The other problem consists of determining the system of
currents on the radiators mounted on the conformal surface when the radiators are
excited via a certain feed network. They key quantities for these calculations are
the scattering coefficients between elements (Section 11.2.2) whose evaluation
requires the inversion of a matrix of a large dimensionality with complex elements.
This is a difficult numerical problem and can be circumvented in certain cases by
using appropriate approximate methods (Section 11.2.4 and 11.4).
In general, the methods of conformal array electromagnetic analysis can be
broadly categorized into two classes:
(a) modal methods
(b) high-frequency asymptotic (ray) methods.
Method (a) is rigorous. However, its area of applicability is restricted to certain
canonical structures having a high degree of symmetry. Nevertheless, the method
is very valuable because the structures that can be analysed exactly are by no
means trivially simple and may in fact constitute accurate analytical models for
practical antennas. For example, the cylindrical periodic array of apertures is a
complex structure that can be studied rigorously (Section 11.4). Moreover, in other
cases, a structure that may be too complex to be modelled 'globally' by using a
canonical geometry can, however, be adequately represented by it in a limited
region of interest.
Approach (b) is a high-frequency method. The radiation of each array element
is represented through a ray system. The method is useful to solve complex prob-
lems not amenable to modal analysis, or for which modal analysis leads to difficult
convergence problems, as, for example, with conical arrays.

11,2.2 Formal characterisation of an array


The current distribution on the aperture of an array, planar or conformal, is not
determined uniquely by the feed network structure but it depends also on the
element interactions. The recognition of the fundamental importance of this fact is
perhaps the most important contribution to the antenna discipline provided by the
modern phased-array theory.8 In Sections 11.2.2 and 11.2.3, a methodology for
the systematic analysis of these effects is established with the aim of providing the
analytical tools for the prediction of antenna patterns, including polarisation and
absolute gain level, and element input impedances.
With reference to Fig. 11.3, an array having an arbitrary geometry and physical
Conformal arrays 231
structure can be modelled as two cascaded networks: the feed network (FN) and
the radiation network (RN). The former performs a complex power division, rout-
ing the input power to the element input terminals with the amplitudes and phases
sought. In general, the FN may have several input ports ('beam ports') to form
independent beams, such as the 'sum' and 'difference' beams of a monopulse sys-
tem, although in Fig. 11.3 a single beam system is shown. Conversely in receive
operation the feed system forms the receive voltages at the beam ports by combin-
ing the voltages at the element terminals, different linear combinations correspond-
ing to different beams. The FN, as defined here, includes also, for phased arrays, the
variable phase shifters. There is, of course, some arbitrariness in the definitions of
the element 'terminals' where the FN and the RN are connected together. For aper-
ture elements, the terminal locations can be conveniently chosen at the array aper-
ture (Section 11.2.4). The RN represents the element external coupling properties
and is characterised by the scattering coefficients between the element terminals.

22 A
o
- T_ ~- B

Fig. 11.3 Network characterisation of an array


{ } : External couplings
fjJ: Internal couplings
* = {a,-}: Free excitation
a radiation network
b feed network

Without any loss of generality, each radiator is assumed to have only one input
port. In fact, if the elements have two excited ports (for polarisation diversity or
polarisation control, for example) the situation will be accommodated simply by
232 Conformal arrays

considering the array as having twice as many elements, geometrically coincident,


each of them having a single port. The transmission lines, or waveguides constitut-
ing the FN, will be assumed to support only either a single propagating mode, or
two degenerate propagating modes for doubly symmetrical cross-sections (circular
or square waveguides).
The 'free excitation' of the array is, by definition, the set of incident waves
pertaining to the fundamental excited mode at the element ports, which will be
considered the components of a numerical vector
{«,}=« (i = 1,2,...,N) (11.1)
having a dimensionality equal to the number TV of the array elements. From Fig.
11.3, it is apparent that, in general, the set of waves {#,-} depends not only upon the
power division realised by the feed network, but also upon the combined effect of
the external and internal element couplings, as discussed later.
A convenient and widely used way of characterising completely the radiative
and impedance properties of an array of radiating elements is based on the use of
the following quantities:
(a) The element 'realized gain pattern' (RGP). This is the array radiation pattern,
including phase, polarisation and absolute gain level, when a single element is
excited by a power generator, the other elements being match terminated. Notice
that in this definition the array is considered, disconnected from its feed network,
and, therefore, only the 'external' electromagnetic interactions affect the element
RGP.
(b) The scattering coefficients among the element input ports due to the external
electromagnetic interactions, called 'external mutual couplings', or simply 'mutual
couplings', in phased-array terminology.
(c) The 'internal mutual couplings', i.e. the scattering coefficients between the
element terminals of the feed system, characterised as a linear network having a
number of ports equal to the number of elements plus the number of 'beam ports',
the latter being the input ports corresponding to different antenna beams.
In accordance with the definition given above, the RGP of the ith element is
evidently equal to the array pattern when the vector a has all the components equal
to zero except the zth, which is equal to unity. Once the set of the RGPs for all the
N array elements has been evaluated, the array pattern for an arbitrary free excit-
ation (i.e., for a certain vectors) is completely determined. Let
?(0, 0) - 04(0, 0) + BgW, *) 01 -2)
represent RGP of the zth element, 0 and 0 being unit vectors in the 0 and0 direc-
tions of a spherical coordinate system. The 0 and 9 components of the RGPs are
complex functions and the phase terms due to the geometric locations of the
elements are incorporated in the expressions of the RGPs. By applying super-
position the array pattern for the free excitation a, which will be called the array
RGP for this particular free excitation, is then found to be
Conformal arrays 233

F(6,d>) = £ ^ ( 0 , 0 ) (11.3)

The array RGP is expressed in eqn. 11.3 via the element RGPs and the array free
excitation.
The element RGP is a particularly useful concept for large cylindrical (and, of
course, planar) arrays with elements on a regular grid.4 In fact, in this case, the
RGPs are identical - a coordinate rotation apart - except for those atypical
elements located near the array edge. In more general cases, the evaluation of the
element RGPs may not be a convenient intermediate step for the calculation of
the RGP of the array. An alternative characterisation of the array properties
employs, along with (b) and (c) above:
(al) the magnetic or electric currents on the array elements when all the elements
are excited by the array free excitation. This characterization is completely
equivalent to that using (a) since, in both (a) and (al), the internal and external
properties of the structure have been taken into account. Once the element currents
have been calculated, the problem is reduced to the evaluation of the radiation of a
known current in the presence of the array surface.
The procedure for the determination of the magnetic currents on aperture
elements will be developed in its general conceptual lines in Sections 11.2.3, 11.2.4
and 11.2.5, the details pertaining to specific geometries being discussed in Sections
11.3 to 11.6.

11,23 In ternal and ex ternal mu tual couplings


Paralleling eqn. 11.1, the set of reflected waves at the element ports is similarly
represented by the vector
{bt} = b (i = 1 , 2 , . . . , N ) (11.4)
The components of a and b are, of course, complex numbers and are normalised, as
customary, so that |tfz-|2 and \bf\2 are numerically equal to the powers associated
with the incident and reflected waves at the ith port. The relationship between b
and a is, in vector-matrix form:
b = S°a (11.5)
where the N x N symmetric matrix S° summarises the scattering (or mutual coup-
ling) properties of the array, as viewed at the element terminals. An element s% of
S° is numerically equal to the complex amplitude of the reflected wave at the port /
when the port k only is excited by an incident wave ak carrying the power of 1 W.
The crucial question of how to determine, through the electromagnetic analysis of
the structure, the matrix S° for each case of interest is postponed to later Sections,
keeping therefore the discussion here on a formal level. Of interest particularly for
high-power transmit arrays is the set of element input reflection coefficients,
defined for an arbitrary free excitation a as the ratios
234 Conformal arrays

r, = ^ (n.6)
at
and called 'active reflection coefficient' at the /th element.
Suppose that 'looking into the network', the elements ports are not matched
and/or the elements are not uncoupled in the sense that the energy fed into one
element is not routed only to the array input terminal but is coupled also to the
other elements and reflected back to the excited element itself. These 'internal'
couplings due to the structure of the FN are graphically indicated in Fig. 11.3 and
and denoted by S-^. It is apparent that if the internal couplings are different from
zero, the free excitation a depends not only upon the transmission set of
coefficients Tt between the array beam ports and the element ports but also upon
the reflected waves b. In other words, if the aperture is not perfectly matched, part
of the energy associated with a is reflected back into the feed network, and, if
there exists internal coupling, back into the aperture again. The process iterates and
multiple reflections occur, affecting the amplitude and phase distributions, that
become very frequency sensitive. Denote by A the excitation at one of the beam
ports of the FN, the other beam ports being match terminated. Simply because of
the linearity of the network the following relationship holds:
a = SNb + TA (11.7)
where T is an N component vector and SN is the iVth-order square matrix represent-
ing the internal couplings between the elements. The physical meaning of T and SJV
is the following. The /th component Tf of T is the normalised voltage at the
terminal of the /th element, when the elements are removed and replaced by
matched loads. An element S^ of SJV is similarly the complex amplitude of a wave
incident from within the network to the input terminal of the kth element when
the elements have been removed and replaced by matched loads and the beam ports
have been match terminated, whereas the /th element has been replaced by a power
generator of unitary power. By using eqn. 11.5, b is eliminated from eqn. 11.7,
obtaining thus:
a = (I^-SJVS0)-1^ (11.8)
where lN is the unit matrix of order N. If the network has no internal coupling, i.e.
if
$N=0 (11.9)
then
a = TA (11.10)
and, in such a case, the free excitation a does not depend upon S°, i.e. it is not
affected by the element mutual couplings. If instead SN is different from zero, the
antenna illumination becomes very frequency sensitive, because so is SN in practical
cases. A physical insight of the phenomena is obtained by resorting to the following
expansion:
2
+ -.- (11.11)
Conform aI arrays 235

which, inserted in eqn. 11.8, yields:


a = TA + SNS°TA + (SNS°)2TA+... (11.12)
having a simple interpretation in terms of multiple interactions and reflections. The
first term of eqn. 11.12 in fact is clearly the free excitation that would be obtained
in the absence of internal couplings. The second term is the modification of the free
excitation due to the energy reflected back into the feed network (because of the
element external mutual couplings) and then back again to the aperture. Higher-
order terms of eqn. 11.12 have a similar interpretation and account for higher-order
multiple reflections, Practically the second and further terms of eqn. 11.12 are to
be considered as distortions of the array illumination sought, affecting the accuracy
of the antenna pattern synthesis.
If at every level of power division in the FN, matched four-port power dividers
are used, namely, matched hybrid junctions with the fourth ports terminated, the
matrix SJV should ideally be equal to the zero matrix. It is, therefore, good practice
to use this kind of 'resistive' power dividers rather than 'reactive' (three port)
devices. In any case, the evaluation of the illumination distortion due to internal
coupling, performed via eqn. 11.12, does not require any matrix inversion. Also,
many elements of Sjy are negligible for practical FN structure; thus, the matrix
$N is sparse, and the computer storage requirement for the computation of SJVS0
is modest.
We proceed now to the discussion of a general procedure for the determination
of the external mutual couplings and the current distributions for arrays of aper-
tures on conducting surfaces. The determination of the internal couplings, in most
cases a straightforward network problem, will not be addressed here.

11.2.4 Multimode aperture elements: Generalised scattering coefficients


Although it would not be too difficult to formally phrase the problem in somewhat
more general terms, the discussion is hereafter conducted in terms of a particular,
although practically very important, type of radiator. It will be apparent that the
methodology of the approach can, in principle, be applied to other more complex
types of elements, although the mathematical difficulty may make the numerical
solution either impossible or impractical.
Consider elements consisting of waveguides, possible filled with a dielectric,
terminated on a perfectly conducting surface, whose radii of curvature at the
element locations are assumed large in terms of the wavelength of operation. The
transition from the feed network will not be in general perfectly matched to the
element waveguide. However, its location will be assumed to be far from the
aperture to make negligible the interaction with the exponentially attenuated
higher-order modes below cutoff which are passively excited at the aperture dis-
continuity. The waveguide cross-section and the dielectric constant of the element
are assumed such that only the fundamental modes propagate. Thus, every higher-
order mode is terminated in its waveguide modal admittance looking into the
element waveguides from the array surface. If the waveguide cross-section is square
236 Conformal arrays

or circular, it will be assumed that no coupling exists between fundamental


degenerate modes due to the element structure. Also, it will be assumed that for the
cross-polarised fundamental mode, the element can be simply characterised by an
admittance term. This assumption is valid in most cases, as, for example, for the
magnetic coupling schematically indicated in Fig. 11.4. Notice that a mismatch in
the transitions from the element waveguides to the feed system for the fundamental
excited mode affects the elements of the matrix $N of the internal couplings,
defined in Section 11.2.3.

Ground
TEn
plane
nominal pol.

Fig. 1 1 A Coaxial to waveguide magnetic coupling. No coupling exists with cross-polarised


fundamental mode
a Cross section
b Electric field

Assume that the first M waveguide modes (ordered with nondecreasing cutoff
frequencies) are sufficient to represent with good accuracy the transverse electric-
field distribution on the element apertures. The choice of the number of modes is
made on the basis of physical considerations and can be validated through con-
vergence tests; i.e. by repeating the calculations of the quantities of interest with an
increasing number of modes and stopping when the results become insensitive to
any further increases.
The superscript k will be used to identify the various aperture modes. The index
k runs from 0 to M— 1, the value 0 being used for the fundamental excited mode.
The modal voltages for the fundamental excited modes are the sum of a set of
incident and reflected waves:
1/2
°7
F°7 = a
= (11.13)
YQ being the characteristic impedance for the waveguide fundamental mode.
With a parallel notation the voltages V* for the passively excited modes can be
represented by the set of vectors, having a dimensionality equal to the number of
elements N:
Vk = {¥/*}, (k = O , . . . , M - 1 , i = 1,...,AO (11.14)
When the array is free excited by the set of waves represented by the vector a, the
following relationships hold, because of linearity:
Con formal arrays 237
1/2 k
Y0- V = 1) (11.15)
fe
6Ofe being a quantity equal to unity for k = 0 and zero for k ^ 0 and S being square
matrices of Nth order. The modal voltages at each element aperture, i.e. the mag-
netic current distribution on the array surface, can be evaluated, (under the Mmode
approximation), for an arbitrary free excitation if the elements of the matrices
Sk, called generalised scattering coefficients, are known. The generalised scattering
coefficients, in turn, can be evaluated from the mutual admittances between the
modal components of the aperture magnetic currents as discussed hereafter.

radiation network

mode '0
element
P mode *1

mode *2

mode' 0'
element
mode ' 1'
q

mode ' 2 '

Fig. 11.5 External couplings. Generalised scattering coefficients

With reference to Fig. 11.5 for k ^0 the meaning of an element Spq of the
N xN matrix Sfe is the following: If the element labelled here by the subscript V is
excited by a free excitation ap - 1, then the modal voltage for the kth mode for the
element labelled 'p' is SpqY0~1/2ap. An intuitive appreciation of the nature of the
mutual coupling phenomena can be possibly achieved by thinking of the radiation
network in Fig. 11.3 as a network having N xM ports, rather than only N, withiV
ports connected to the feed network and (M — 1) x N ports terminated by either
modal admittances (for higher-order modes), or by an admittance determined by
the structure of the transition from the feed network to the element waveguide
(for the cross-polarised fundamental mode), Fig. 11.5.
238 Conformal arrays

11.2.5 Mutual admittances and scattering coefficients


The evaluation of the scattering coefficients for the fundamental mode is essential
for determining the array free excitation a (eqn. 11.15). The generalised scattering
coefficients for higher-order modes are necessary, on the other hand, to evaluate
the modal content of each aperture magnetic current (eqn. 11.15). A procedure
for the evaluation of the generalised scattering coefficients is now given.
Introduce a position vector s, identifying a generic point on the array surface.
Let e%(s) be a vector function representing the electric transverse field distribution
of the kth mode on the pth aperture. The modal functions are normalised as
follows:

JJ (11.16)

the integral being extended over the pth aperture. Let hlq(s) be the transverse mag-
netic field on the array surface generated by a magnetic-current distribution on the
qth aperture. The magnetic current is assumed to have the functional form corre-
sponding to the i th mode with unit modal voltage, that is, having the form elq(s) x ft
with h the normal to the array at the location of the qth aperture. In evaluating the
magnetic field, the apertures are considered absent; i.e. the metallic continuity of
the array surface is imagined to be reconstituted. The admittance between the
modes k and i (possibly equal) on the apertures p and q (possibly coincident) is
defined by the expression:

(11.17)

For a fixed pair of modal indexes 4fc' and H\ the quantities YpQ form a matrix of
order N denoted by Ykl. The following matrix of M x N rows by M x TV columns
can then be formed:
Y oo Y01 ... Y0^"1
(11.18)
yM-1,0 rM-\,M-\

To proceed in relating the mutual admittances to the scattering coefficients, intro-


duce the set ofM diagonal matrices of order N:
0 0
Yk 0
(11.19)

- 0 0 Yh J
w
The diagonal elements of D are all equal to the modal admittance viewed by the
&th mode 'looking into' the element (Appendix 11.8.1). Form the matrix
Con forma/ arrays 239
r D (o) 0
0 0
D = (11.20)

- 0

and finally consider the square matrix of order N x M:


S = (D + Y r ^ D - Y ) (11.21)
whose evaluation requires the inversion of a matrix also of order NxM. It is
possible to show that the matrices Sfe defined by eqn. 11.15 are submatrices of the
matrix S. To elaborate, it can be shown that the following partitioning holds:

s1
s = s2 (11.22)

where the tilde (~) indicates transpose and the broken lines stand for a matrix of
order ( M - l ) x J V .
Direct inversion of the matrix in eqn. 11.21 is difficult except for very small
arrays. Assuming, for example, a number of elements N = 300 and a number of
modes M = 3, a relatively simple characterisation of the element magnetic currents,
then in order to identify the submatrices Sfe of S, the inversion of a complex matrix
of order N xM= 900 is necessary. This is a difficult and time-consuming task even
with present high-speed computers.
The problem can, however, be circumvented by using a series-expansion method
for the evaluation of eqn. 11.21. Consider the identity:

(D + Y)"1 = Ifl-Kl-D^ (11.23)

where 1 is the unit matrix of order M x N. Assume that the norm of the matrix in
the inner parentheses in eqn. 11.23 is 'small' in the following sense:
-D"1Y)|i< 1 (11.24)
It can be shown that eqn. 11.24 holds in any practical case, except for large
element mismatch. Then the expansion

(11.25)
is valid and when inserted in eqn. 11.21, taking into account eqn. 11.23, gives:

S = (11.26)
240 Con forma I arra ys

not requiring any matrix inversion. For cylindrical arrays having a large number of
elements arranged in a regular grid, it is possible to avoid large matrix inversions by
resorting to a different approach that uses the high degree of symmetry of the
structure and neglects array edge effects. The philosophy is similar to that widely
used for planar array analysis. A discussion and numerical examples are given in
Section 11.4.

11.3 Cylindrical array of apertures

In Section 11.2.5 the formal expressions for the generalised mutual admittances for
aperture elements were given. A procedure was discussed to obtain the element
mutual couplings and the magnetic current distribution on the array surface under
the assumption that M aperture modes represent adequately the radiator transverse
electric-field distributions. This formalism will be applied in this Section to the
study of an array of apertures on a conducting circular cylinder. To provide the
necessary background, certain well known results concerning the field of a magnetic
current flowing on a cylindrical conducting surface will be briefly recalled. The
cylinder is assumed infinitely long, that is no edge diffraction effects are here
considered.

11.3.1 Field of a magnetic current on a circular conducting cylinder


With reference to Fig. 11.6, the axis of the cylinder is assumed to be coincident
with the z axis of a rectangular system of coordinates. A polar system, r, 6, 0, and a
cylindrical system p, $ and z are associated in the usual way with the rectangular
system. In addition to the angular coordinate 0, it is convenient to introduce a
curvilinear coordinate on the cylinder surface defined as follows:
% = R(j> (11.27)
R being the cylinder radius. Thus, a point on the cylinder surface can be defined by
the position vector:
s = (£,z) (11.28)
Consider an aperture, whose geometric centre is located at z = <j> = 0. Let the trans-
verse electric-field distribution on it — a 2-dimensional vector — be denoted by:
e&z) = * 0 ( £ , z ) 0 + *,(£,*)£ 01.29)
Introduce the Fourier transforms (FT) of the circumferential and axial components

w) = ± jjA e&, z)e^+wz)d$ dz (11.30)

\\ez{hz)ed$dz (11.31)
J J
2TT
Conformal arrays 241
the integrals being extended to the aperture, and u and w being the variables conju-
gate to £, z in the FT. Eqns. 11.30 and 11.31 can be lumped together and written
concisely as follows:

l
'*dA (11.32)

aperture p

Fig. 11.6 Cylindrical array geometric reference

where

u=(u,w) (11.33)
is a position vector in the u, w plane and
E(W) = e^{u)4>^ez{u)z (11.34)
The procedure for the determination of the field in the region p>R is
242 Conformal arrays

straightforward.10 Assuming a time dependence exp/cof, with the present notation


the expressions for the electric field are promptly found to be

and

(11.36)
where & is the free space propagation constant and
£ = y?k2-w2 (11.37)
with the branch of the square root chosen to satisfy radiation conditions, i.e.
\\/k2-~w2\ for \w\<k
, o (1138)
~j\\/k2 -w2\ for |iv|>fc
) and /5/^2)'( ) are the Hankel function of order n and second kind and its
derivative, respectively.
Expressions for the tangential magnetic field on the cylinder surface p = R are
similarly established. By using the coordinates £, z,

xM^w| + :^ e J^w|| (H.39)


and

(1L40)
jk H^(R^)}
Y^ In \\

which identify the currents on the conducting surface of the cylinder, T? denotes the
characteristic impedance of free space.

113.2 Harmonic series expressions for the mu tual and self-admittances


Because of the symmetry of the circular cylinder, the mutual admittances between
the modes of two apertures 'p' and V depend only upon the relative positions of
the two apertures, i.e. upon
Conformal arrays 243

spg = sQ-sp (11.41)


sp and sq being the position vectors identifying the two aperture locations, accord-
ing to eqn. 11.28. Because of eqn. 11.41, eqn. 11.17 becomes

y%Pq) = j \ A ek(s) x h\s ~spq)'hdA (11.42)

the integration being over the aperture of the reference element whose centre is at
z = (j) = 0. More explicitly by introducing the circumferential and axial components
of the field:

Yki(spq) =jjA [4(*)Ai(s-s M )-€*(z)^(s- SiHr )]d{dr (11.43)

Introduce for the FT of the transverse component of the fcth mode the following
notation, in accordance with eqn. 11.32:

e\u) = — ff ek(s)eiu'sdA (11.44)


2TT JJA

Explicit expressions of the FTs of the mode functions for rectangular and circular
waveguides are given in Appendix 11.8.2. Introduce in eqn. 11.43 the expressions
11.39 and 11.40 for the transverse components of the magnetic field. The
expression for the generalised mutual admittance becomes;

\ )\~e ^'

Rt;Hpm)6'[R'W|

xexp[-j^tpq -jwzpq)- (11.45)

The expression 11.45 can be evaluated without further analytical manipulations,


performing the integration with respect to w numerically. The azimuthal series is
slowly convergent, however, and should be truncated to a number of terms
n>$R (11.46)
This requirement makes the evaluation of eqn. 11.45 impractical for arrays on large
cylinders. For large arrays on large cylinders the periodic array approximation leads
to a simpler evaluation of element interactions and array pattern as discussed later,
see Section 11.4. For an array on a cylinder with a large radius having small or
moderate number of elements, the amount of computation can be reduced by
244 Conformal arrays

using eqn. 11.45 for the generalised self admittance only (sPQ = 0), and using
instead in eqn. 11.43 high-frequency asymptotic expressions for the magnetic field
as discussed in later Sections. For the self admittance a further simplification con-
sists of replacing the ratio of the Hankel function and its derivative with its asymp-
totic value for large arguments:

(1L47)

which amounts to replacing locally the cylindrical with planar geometry.

11.3.3 GTD surface field calculation


It is possible to use the Watson transformation23 — or the equivalent procedure dis-
cussed in Section 11.3.5 in connection with the radiation pattern evaluation — to
transform the series (eqn. 11.45) into a rapidly convergent series. A detailed dis-
cussion on the related problem of the calculation of the current on a cylinder
induced by a rectangular slot is in Hasserjian and Ishimaru.11 Here the less rigorous
but simpler approach is taken of postulating a GTD Green's function. The
formulation here discussed is largely based on the work by Lee.15
A ray (or GTD) expression for the magnetic field to be used in eqn. 11.43 is
sought. To proceed, consider two generic points s and s' on the cylinder surface.
Clearly, the points can be reached from s' via an infinite number of geodetic helical
paths on the surface. However, rays which travel more than once around the
cylinder will be neglected, Thus, only the two rays from s to s surrounding the
cylinder clockwise and counter-clockwise along the shortest paths will be con-
sidered. Also, if the lengths of the two paths are widely different, only one of them
will contribute significantly to the field. In the £, z reference, the components of s
are £' and z . The components of t ($', s) are (Fig. 11.7):

(1L4S)
" = i^7i

The unit vector in the direction normal to the surface at the points is denoted by
h(s). The binormal to the geodetic ray at s is defined by
i(s-s')xh(s) = b(s-s) (11.50)
whose components are
b^s-s) = tM(s-s) (11.51)
bz(s-s) = -rt(s-s'). (11.52)
Introduce two functions Hb(s) and# t (s), where
Con forma/ arrays 245

(11.53)
The expressions for Hb(s) and Ht(s) will be given later. Their physical meaning is
that of the components of the GTD dyadic Green's function for an element of mag-
netic current tangential to the cylinder.14 The tangential magnetic field on the
cylinder surface excited by the /th mode on the reference aperture is then given by:

*'(*) =\jA, {b(s-s')Hb(\s-s'\)[b(s-s)'et(s')xh\

+ i(s-s')Ht(\s-s'\)[i(s-s') «]} (11.54)

Fig. 11.7 Field at s is the sum of two ray contributions propagating in the directions t and tl

I' and z being the coordinates of a generic point of the reference aperture. The
radius of curvature is constant on the ray geodetic path and is equal to
R
Rt = (11.55)
cos2 5
where 6 is the angle formed by i and the % axis. The radius of curvature in the direc-
tion of 5 is also constant and equal to
R
(11.56)
sin2 6
246 Conforms/arrays

Introduce a distance parameter

/ * \1/3
7 = s (1L57)
(ai?)
where k is the free-space propagation constant. Introduce the function
k2 e~iks
G(s) = — (11.58)

and the Fock functions9 u{ ), p( ), w'( ), v'{ ) whose expressions are given in
Appendix 11.8.3. Then the longitudinal and transverse GTD Green's functions for
8 ¥=n/2 are given by:

( |) J (11.59)

+ j{yf2kRty2f* tan2 Se/(r) C(s) (11.60)


The numerical evaluation of the Fock functions is addressed in Appendix 11.8.3.
If the radius of the cylinder becomes infinite, from, eqns. 11.59 and 11.60,

- \I-J- (41
[ ks \ksf J
(11.62)

which are identical to the exact expressions of the b and i components of the mag-
netic field of an elementary magnetic dipole on a ground plane having the orient-
ation of b and i, respectively. For 6 -> TT/2

•is) (11-63)

which, in terms of the solution for a ground plane, is written

^9 i -64)
It can be shown easily from eqn. 11.64 that in the axial direction the field on the
cylindrical surface is stronger than on a plane. For very large r, i.e. when the
observation point is in the deep shadow,
Con formal arrays 247

Ht(s)~-j-Hb(s) (11.66)

In the deep shadow the field is a slow wave decaying exponentially along the sur-
face ray.

11.3.4 GTD expressions for the mutual admittances


Clearly, the magnetic field excited by the ith mode on the aperture centred ats PQ is
obtained by replacing s with

in eqn. 11.54. Since there is clearly no reason to use the double subscript 'p, q\ it
will be replaced from now on by the single subscript ' p \ By using the fact that
f, h9 b form a normal trihedron, from eqns. 11.43 and eqns. 11.51-11.54 one
obtains for the GTD equivalent of eqn. 11.45:

Yk%) =- j | A d%dzek{s)'\\A,d%tdzti{s-sp-st)Hh{\s-sp-sf\)

d%'dV''b{s-sp-s')

01-67)
which is the generalised mutual admittance for the ith and fcth modes for two aper-
tures whose relative position is identified by sp.

Example: Two equal rectangular circumferential slots. Mutual admittance


between fundamental TE 10 modes axially polarised:
Circumferential dimension = 2a
Axial dimension = 2b
The components of i(s ~sp — s') are:

z
Is —Sp —sV is — sp — s\
where

and the components of b(s — spq — s) are given via eqns. 11.51-11.52. For the
fundamental mode the direction of the electric field on the aperture is constant.
Thus, for axially polarised modes,

' ~ ' (11.70)


248 Conformal arrays
and
b(s-sp-s')-e°(s) =
1
cos V2 (11.71)
u- ,-s'l yjlab
The other scalar product appearing in eqn. 11.67 is expressed in a similar way. The
final expression is:

xcos-- (11.72)

In a parallel way, the mutual admittance between two axial slots can be established.
The straightforward details are here omitted. In Figs. 11.8-11.10 numerical
examples are given,* and also a comparison is made between the results obtained via

160

Fig. Mutual admittance Y00 between the fundamental TEl0 modes for two circum-
ferential slots as a function of the azimuthal separation <f>0. Vertical separation

mag.
GTD solution
phase
o exact modal

GTD and modal expansions. Notice that the phase in Fig. 11.8-11.11 is normalised
by multiplying Y00 by the geometric phase term exp (jksp).
The explicit expression of eqn. 11.67 is not as simple as eqn. 11.72 for modes

S. W. LEE. Private communication.


Con formal arrays 249

having the polarisation of the electric-field variable over the aperture. This occurs
for all modes of circular waveguides and for some higher-order modes of rectangular
waveguides. However, the numerical evaluation is straightforward.

-60 85°

05
o
-70 80° £

-80 75° E

-90 70°

-100 JJ65°

z o /A
m
Fig. 11.9 Mutual admittance Y between the fundamental TEl0 modes for two circum-
ferential slots as a function of the vertical separation z0. Azimuthal separation

GTD solution mag


— phase
o exact modal

For large interelement spacing a drastic simplification of the GTD expressions


for the mutual admittances can be established. If the aperture maximum linear
dimensions 2a ^spy then for any points ands' of the first and second aperture,
sf- sp) ~ip = (cos So, sin 5 0 ) (11.73)
where So is the angle the vector sp forms with the % direction (Fig. 11.11). Similarly,
6 ( 5 - 5 ' ~~sp) ~hp = (sin 5 0 , - cos So) (11.74)
To proceed, consider each of the functions Hb(s) and Ht(s) as the product of a
slowly and a fast varying function of s:
Hb(s) = P(s)e-ihs (11.75)
k
f*t(s) = (11.76)
where a comparison of eqns. 11.75 and 11.76 with eqns. 11.59 and 11.60 identifies
the slowly varying functions P(s) and Q(s). The simplification proposed by
Steiskal21 - although in connection with a different GTD Green function -
consists of replacing in Q(s) and P(s) their argument with sp. For the phase term the
following 'far zone' approximation is used (Fig. 11.11):
250 Conformal arrays

\s-s -sp\~sp + ip'(s-s') (11.77)

Thus, eqn. 11.67 reduces to the approximate expression

Yki(sp) - -

z\\bp-jj (11.78)

-70! -100°

-90 -120°
•o

-110

-130 -160°

-150 -I ~Q -180°

Fig. 11.10 Mutual admittance Y00 between the two fundamental TElQ modes for two axial
slots as a function of the vertical separation z 0 . Azimuthal separation 0O = 0
— mag.
GTD solution
phase
Exact model o

By introducing the FT of the mode functions, eqn. 11.44, expr. 11.78 becomes:

-Q(sp)Uh(-kip)*bp][bp' t\ktp)]} (11.79)


which does not require any integration for its numerical evaluations. As expected, if
in eqn. 11.79 the asymptotic expressions of P(s) and Q(s) for very large cylinder
radii are used (eqns. 11.61 and 11.62), then eqn. 11.79 becomes identical to the
asymptotic expression for large separation of the mutual admittance between two
single-mode apertures on a ground plane.3 As an example of application of eqn.
11.79, consider the mutual admittance between the modes T E n (circumferentially
Con forma/ arrays 251

polarised) and the mode TMOi on two apertures whose relative location is indicated
again by sp. In the expressions for the FT of the mode functions (Appendix 11.8.2),
according to eqn. 11.79, the following replacements are made:
t<-k (11.80)
and
H<-80 (11.81)
for the second aperture, and
(11.82)
for the first aperture. Also, introducing the polar components of the FT of the
vector mode functions, the following identity is recognised:

= e,(*,8o) (11.83)
and
t{kip)'b = (11.84)

aperture q

aperture p

Fig. 11.11 Geometric relationships on the developed cylinder

Consequently, from (11.79) and the results of Appendix 11.8.2.

(11.85)
252 Conformal arrays

valid for large separation sp. In the above, the roots xnp and xnp of the Bessel func-
tions are as defined in eqns. 11.221 and 11.223.

11.3.5 Radiation pattern evaluation


Once the modal voltages for the apertures have been calculated, as discussed in
Sections 11.2.4, 11.3.2 and 11.3.4, the transverse electric-field distribution on the
cylindrical surface (in the M mode approximation) is given by the expression

Etis) = 1 1 Vfeh(s-sp) (11.86)

where it is recalled that N and M are the number of array elements and aperture
modes, respectively, and sp = 0 identifies the reference element located at
z = (p = 0. The FT of eqn. 11.86 is (see eqn. 11.32):

t(u,w) = £ *p(u,w)eju'sP (11.87)

where

•P(«,HO = ± ff £ Vpkek(s)eiu'sdHdz (11.88)

Insert the z and <p components of eqns. 11.35 and 11.36 into eqn. 11.87. Introduce
the asymptotic expression for the Hankel function for large arguments. Thus for
large p the pattern for the array radiated field is:

dw (11.89)
and
EJp,<p, z ) ~

(11.90)
Expressing f in terms of w by using eqn. 11.38 the integrals in eqns. 11.89 and
11.90 can be evaluated asymptotically (for large z) through a standard application
of the saddle-point method 4 . The following expressions are thereby obtained for
the 0 and $ components:
Conformal arrays 253

and, noting that Ee-+— Ez/sin 9;

e'lkr 1 -

(1L92)
F)
The numerical evaluation of eqns. 11.91 and 11.92 is straightforward. The number
of terms required is approximately equal to and greater than the argument of the
Hankel function
\n\>~kRsin6 (11.93)
The series are therefore slowly convergent and of impractical use for large cylinder
radii. A procedure for transforming eqns. 11.91 and 11.92 in computationally
simpler expressions is presented below. The key idea is the transformation into a
rapidly convergent series of which only one or a few terms are retained. The
procedure is different but equivalent to the Watson transformation used in scatter-
ing problems23 (for an introductory discussion see Reference 6).
To introduce the technique consider a continuous function C(v) of the argument
p. The value taken by C(y) for v equal to an integer n must be equal to C(n). By
using the sifting property of the Dirac delta functions, the following identity is
established:

lL exp \~in ( 0 ~f)] c ( w ) = C e x p f ~jv ( 0 ~f


+ OO

x £ b{v~ri)dv (11.94)

Because of the well known distributional identity

+
f 8(v-n) = £ ei2ww (11.95)

from eqn. 11.94

f C(n)exp\-jnL-Z-

fjexp - ; > U - 2 7 r « - ^ j c ( I ' ) ^ (11.96)

which is the well known Poisson summation formula


254 Conformal arrays

The identity (eqn. 11.96) will be used for the evaluation of the pattern com-
ponents (eqns. 11.91 and 11.92). Different mathematical expressions will be given
for apertures in the lit and shadow regions with respect to the directions of
observation. The patterns of the whole array will be calculated here as the super-
position of the patterns of the single apertures which can be either in the lit or in
the shadow region. Use the superscript *p' for the pattern of the generic pth aper-
ture. By using the identity eqn. 11.96 in eqns. 11.91 and 11.92 (here applied to a
single element) the following expressions are obtained, after allowing for the
appropriate phase terms due to the element location:

irR n

1
v \ . v cos 6
P ,k^8j+jejB,,ifcco,0|| dv (H.97)
I \
and

£ . \l (v , J ^
2"
(11.98)
which is the starting point for the derivation of relatively simple expressions valid
for large kR sin 6.

..O—J o O 1 O 1 O 1 O O
-3-rr 4>_) -2tT -TT 0O 4> rr 0, 2TT 3TT

2Tr 2TT
I*. 2TT ^r •*«

Fig. 11.12 Extended angular space


4> — direction of observation
source location

Expressions 11.97 and 11.98 admit an intuitively simple interpretation. Each


integral may be thought as the contribution of a source at the azimuthal location
<j)n — UplR) — 2nn in an extended angular space going from — °° to 4- ©o. The situ-
ation is graphically illustrated in Fig. 11.12. The position of the physical source, in
the physical space — extending from — n to + n — is indicated by 0O — Ip/^- It is
Conformal arrays 255
intuitively clear that the contributions of the image sources, located at angular
distances from 0o greater than 2TT, contribute little to the radiation pattern and can
be neglected. Mathematically, the reason resides in the highly oscillatory nature of
the integrands in the various terms of eqns. 11.97 and 11.98, that makes the
integrals pertaining to the image sources comparatively negligible.
In Appendix 11.8.4 it is shown, that for a large cylinder, the far field can be
approximated as follows:

(a) Lit region: The minimum azimuthal separation between the aperture
angular location and the direction of observation is denoted by ± |^ 0 |, with the
positive sign if the direction of observation is reached from the aperture moving
counterclockwise along the shortest arc. In the lit region \\po\ <TT/2. An approxi-
mation for the far-field expression is

£$(?> QJ 0) ~ ~~7 k cos \j/0 e x P (Jk [zp cos 6 -f R sin 8 cos ^ 0 ])


r -»e» r
x ezp{k sin 6 sin # 0 , k cos 6) (11.99)

E$(r>6,0) — / —— k exp {jk\zp cos 0 + R sin 6 cos i//0])[sm 0e<j>p(k sin 0

x sin $o> k cos 8) + sin i|/0 c o s 0€ zp(k sin 0 sin ^O, k cos 8)] (11.100)
These expressions are recognised to be the same as the ones that would hold for an
aperture — having the same magnetic-current distribution — on a ground plane
tangent to the cylinder at the element centre.

Fig. 11.13 Azimuthal separation between source and direction of observation


<t> — direction of observation
0O = aperture azimuthal location
256 Conformal arrays
(b) Penumbra and shadow region: Call $* — | $ + | and \jj~ = — \\jj~\ the angles,
counted counterclockwise and clockwise, respectively, between the aperture and
the direction of observation (see Fig. 11.13). The smallest of \jj+ and | ^~| is clearly
the angle 11//01 already defined. In the penumbra and in the shadow regions TT/2 <
|\// 0 |<7r. By introducing the Fock functions f(t), g(t) and g^l\t) defined in
Appendix 11.8.3 the following expressions for the far field are obtained (Appendix
11.8.4):

e'!kr 1 ( f / n
E$(r, 0, 0) ~jk — exp (jkzp cos 0) exp — jkR sin 0 I \jj+ — ~
r~*oo r 2 II \

+
QpQc sin 0, k cos 0) sin 0 4- ezp(k sin 0, k cos 0) cos 6]g j ^ 1

2sin 2 0j

x ezp(k sin 0, k cos 0) 4- exp —jkR sin 0 I I \p~ I — -

x
[ e 0p(~ ^ sin 0, k cos 0) sin 0 — ezp(— k sin 0, k cos 0) cos 0]

x (|*R sin ^) 1 / 3 e^(--fcsin 0, * cos 0) (11.101)

and

x | exp —jkR sin 0 | $ + — -—


J 11
sinsin
6e6e zp(ksinsin
zp(k 0, 0,
& kcos
cos0 0)/1( i|/+——~~\
) / I $*

x QikR sin 0) 1 / 3 | + exp | -jkR sin 0 || ^ |

xsin0e zp (-A:sin0,A:cos0)/\\\4>~\ ~\(h^R sin0) 1 / 3 | J (11.102)

Notice that in eqns. 11.101 and 11.102 the quantity


Con forma/ arrays 257

is the curvature radius for an helical path on the cylinder whose tangent forms an
angle S with the cylinder generator. Introduce the path lengths

01.104)
and according to eqn. 11.57, the distance parameters
1/3

7* = s± (11.105)
2R2t(9)
It is evident that the arguments of the Fock functions in eqns. 11.101 and 11.102
are equal to eqn. 11.105, i.e.

:
--\(\kR sin$)1/3 = T± (11.106)

Expressions 11.101 and 11.102, with the help of eqns. 11.104,11.105 and 11.106,
lead to an interesting and illuminating physical interpretation of the radiation
mechanism. The portion of the electromagnetic energy that is not radiated in the lit
region flows around the cylinder and is continuously radiated tangentially. Neglect-
ing the contributions of surface rays completely encircling the cylinder, to each
direction of observation are associated two diffraction points, whose distances from
the source along the geodetic rays are s+ and s~, respectively. The two surface rays
form an angle 6 with the cylinder generator and their radii of curvature are given by
eqn. 11.103, At the diffraction point the rays leave the surface tangentially and
propagate towards the observation point.
For large cylinders, unless \\}/~ \ ~ \p* ~TT, the contributions pertaining to i//0,
the smallest of the two angles, will be dominant. Thus, only one of the two terms in
eqns. 11.101 and 11.102, representing the two contributions, needs to be
considered.

11.4 Cylindrical arrays with a large number of elements: The periodic structure
approach

11.4.1 Periodic infinite array model


It is well known that the evaluation of the mutual couplings and of the radiation
pattern of a large planar array is drastically simplified if the array edge effects are
neglected. The idea consists of modelling the actual finite array as a double periodic
infinite structure.8 In this idealisation all elements are assumed to experience
identical mutual coupling effects. Thus the effect on the array performance of the
different electromagnetic behaviour of the atypical elements located close to the
array edges is ignored. The approach can be extended to the analysis of large
258 Conformal arrays

cylindrical arrays having the elements arranged on a regular grid. The actual finite
array is considered embedded in an infinite periodic structure. Neglecting the array
edge effects makes all the element RGPs identical, except for a rotation and a geo-
metrical phase term.
The method of analysis is based on the systematic use of the rotational and trans-
lational symmetry of the double periodic structure. Every arbitrary free excitation
of a finite portion of the infinite array is considered as a weighted superposition of
a set of elementary excitations of the infinite array matching the symmetry of the
structure, in the sense of generating fields equal for each array elementary cell,
except for a uniform phase progression. These elementary excitations are eigen-
vectors of the symmetry operators representing the congruences of the structure.16
Since the generalised scattering matrices Sfe defined in Section 11.2.4 commute
with the symmetry operators, the eigenexcitations are also eigenvectors of Sk. The
analysis of the structures for these eigenexcitations is much simpler than for a
generic excitation. The general case, on the other hand, can be analysed by applying
superposition.4 These concepts lead to a relatively simple method of analysis, and
also to an effective technique for element pattern design.
The basic ideas of the method are described in the next Section. Details of
analytical or numerical nature are presented in Sections 11.4.3 and 11.4.4.

0 o
o olo o 0

Fig. 11.14 Periodic infinite array

11A.2 Eigenexcitations and eigenpatterns and their use in array analysis


As mentioned in the previous Section, the large array under study is assumed
embedded in an infinite array having the same element grid. The geometric reference
Conformal arrays 259
has been specified in Section 11.3.1, Fig. 11.6. The double periodicity of the
elements is for the moment left arbitrary: the elements can be arranged either in
a rectangular or in a triangular grid. The number of elements of each array 'ring' is
denoted by L. With reference to Fig. 11.14, the symmetry of the structure is
defined by the two column vectors identifying the element lattice. For example,
for elements in an equilateral triangular grid - as shown in Fig. 11.14 - the two
vectors are

., - [I (H.10T)

s2 = (11.108)

A point of the cylinder surface can be identified by the position vectors, given by
eqn. 11.28. Again the element whose centre is located at £ = z = 0 is called
'reference element'. The position of a generic element is identified by the vector
smn = mSl+m2 (11.109)
where the index n designates a particular ring and m the element location within it.
The matrices Ykl defined in Section 11.2.4 have in the present case an infinite
number of rows and columns. The translational andL-fold rotational symmetry of
the structure is reflected in the fact that the mutual admittances between the two
apertures identified by the pairs of indexes m, n and p, q depends only upon the
differences of the indexes (m—p mod L)
Yik[(m-p)s1+(n-q)s2] = Y*p.n>q (11.110)
Of course, the same property holds for the elements of the generalised scattering
coefficients:
sk[(m-p)sl+(n-q)s2]=s^p,niq (11.111)
On the basis of the special structure of eqns. 11.110 and 11.111, it is easy to show
(Appendix 11.8.5) that the eigenvectors of the matrices whose elements are given by
eqns. 11.110 and 11.111 are the infinite dimensional vectors (suitably normalised)

ex
e(u0, wo) = e(uo) = J^£ P [-/Ho("»i + ns2)] (11.112)

where m is an integer between 0 and/, — 1 and n is any integer. Each vector


«o=["o,wo] (11.113)
identifies a particular eigenexcitation. Express the first component of w0 as follows:

«o = ^ (11.114)
K
260 Conformal arrays

which indicates that the phase progression between two adjacent elements in a ring
is 2mjN. The 'azimuthal number' / can take only integers values ranging from 0 to
N — 1, or from — N to N — 1, for elements in a rectangular or in an equilateral tri-
angular grid, respectively. The longitudinal number vv0 takes all real values from
— rr/h to n/h, and from — 7r/2/z to ir/2h for the rectangular and the equilaterla tri-
angular grid, respectively. The justification of these statements lies in the concept
of a reciprocal cell in the wavenumber plane. In the planar array context a dis-
cussion is found in reference 3.
Because of the very definition of eigenvectors of an operator, the following
equations hold:
Yhie(u0) = Yki(uo)e(uo) (11.115)
and
Ske(u0) = rk(u0)e(u0) (11.116)
with Y (u0) and rk(u0) eigenvalues associated with the eigenvector e(u0). From
ki

eqns. 11.21, 11.115 and 11.116 it follows, via some straightforward algebraic
manipulations, that the eigenvalues rk(u0) must satisfy the set of M linear
equations in M unknowns (M being the number of aperture modes approximating
the element magnetic currents)
M-l
I = -Ys0(u0)+Y08s0,
s0

fe = O

(s = 0 . . -Af—1) (H.117)
which allows one to determine the Tk(u0) if the set of numbers Ykt(u0) is known.
The expressions for the Ykl(u0) are given in the next Section.
In Appendix 11.8.5, it is shown that the system of vectors (eqn. 11.112) form
an orthonormal complete set for u0 belonging to the appropriate limited set of
points in the uOi w0 plane. This means that any arbitrary excitation a of a finite
portion of the infinite array can be represented as a linear superposition of the
vectors e(w0).
To make the analytical development less abstract, from now on the discussion
will focus on the triangular arrangement of elements shown in Figs. 11.14 and
11.16. The minor modifications of the various equations for other element grids
such as the rectangular are easily worked out and are not reported here.
Any free excitation of the finite array embedded in the infinite periodic struc-
ture can be expressed as a weighted superposition of the vectors (eqn. 11.112),
because of their completeness property (Appendix 11.8.5). The following expan-
sion holds:

a
= 1 } nJ2hcii>Wo)e[R>Wo)dWo (11.118)
The coefficients c(i9 w 0 ), because of the orthonormality of the basis vectors (eqn.
11.112) (see Appendix 11.8.5), are obtained by projecting a on eache(/, w0):
Con formal arrays 261

c(i,w0) = e\uo)'a = £ I amn J— exp[ju0- (mSl + ns2)]


m n V 2nL
(11.119)
the '+' denoting the conjugate transpose.
How to make use of eqn. 11.119 to find the modal voltages at the element aper-
tures for a finite array? Reiterating that the edge effects are neglected, the pro-
cedure is the following: The array is imagined to be embedded in an infinite array.
Then apply eqn. 11.15, with a expressed via eqn. 11.118, and use eqn. 11.115,
11.116 and 11.117. This requires the determination of the eigenvalues Yki(uo)9
discussed in the next Section. By using superposition the modal voltages for the
element identified by the indexes >, f are found to be expressed as follows:

1/2 k
Y001/2VVrrt t
Y = art80k+ £*

x {exp/iio-[(ro ~ r)si + (n - t)s2]}amn (11.120)


the summation being extended only to the elements belonging to the actual finite
array.
The radiation pattern can be determined via superposition by using as elementary
components the radiation patterns corresponding to the eigenexcitations which
will be called eigenpatterns. Denote by g(i, w0, r, #, 0) the far field radiated by the
periodic structure when it is free excited by the eigenexcitatione(///?, w 0 ). Then the
radiation pattern F(0, <p) for an arbitrary excitation a of the finite array is given by:
e-jkr ^ L-l rn/2h

— F(0,0) = £ j _ f f / 2 h !(i,wo;r,0,0)c(i,wo)rfwo (11.121)


i = —L
where again eqn. 11.118 has been used. The determination of the eigenpatterns
g(i, w, r, 6, 0) will be discussed in Section 11,4.4 where it will be shown that the
integration in eqn. 11.121 is performed via an application of stationary phase
method and does not require numerical quadratures.

11.4.3 Evaluation of the eigenvalues of the admittance matrices


An expression for the eigenvalues is obtained via a direct calculation of left side of
eqn. 11.115. By using eqn. 11.42, the mutual admittance for the modes k and /
between the apertures located at the point m, n andp, q of the lattice, respectively,
is given by

j\ (11.122)

Thus, thep, q component of Ykle(u0) is


L-l
[Yki(uo)e(uo)]pq ]pq =
= 1 1 Yki[(p-m)Sl+(q-n)S2]

exp [-juoimsx + ns2)] (11.123)


262 Conformal arrays

Inserting eqn. 11.123 into eqn. 11.115

Yki(u0) exp [-fuo' (psx + qs2)] =

L-l +<*>
Z L hl(s — msi — ns2) exp [— juo(msi + ns2)] d%dz (11.124)

To proceed we need an integral representation of the magnetic field h\s) supported


by the /th aperture mode on the reference element. This is done by formally writ-
ing the Fourier series (eqns. 11.39 and 11.40) as Fourier integrals. Put

(1U25)
]
. IU »\ ' [«'"• " S e n

where u represent the pair of wavenumbers u, w. Consider 11.125, 11.126 as


components of a two-dimensional vector
*"(«) = Mrj(u)q + arj(u)z (11.127)
Thus the tangential magnetic field on the cylinder is written

I 6 \u~l\dudw, (11.128)
2TTR

Therefore, by recalling eqn. 11.95,

A'(,-mSl-«s2) = ^- {_
2n
1
(11.129)

which, inserted in eqn. 11.124, and interchanging the orders of the integration and,
recalling eqn. 11.115, gives, through simple manipulations,
1 4- OO + °° •*" ° °

Yki(u0)exp[-ju0-(ps1 + qs2)] = — f + °° I ' I I


Z7T J_o

xexp[/(«— «o)( ms i + j
(11.130)
Con forma/ arrays 263
Notice that now both the summations are extended to infinite limits. Using eqn.
11.44, one obtains

xexp {](u -uo)[(m -p)sx + (n -q)s<i\}h x tk(-u)dudw (11.131)


To proceed we need to evaluate the sum under the integral sign in eqn. 11.131. To
this purpose, paralleling the planar array case,3 introduce the vectors tu h
'reciprocal' to the lattice vectors st,s2. They are defined by the equations:
*CSh = 2n8ik, (Uk = 1,2) (11.132)
For the equiangular triangular grid of Figs. 11.14-11.16 whose direct lattice
vectors are given by eqn. 11.107-11.108, the reciprocal vectors are given by:

f •-;]• ' - [ ° - T ]
It will prove convenient in view of further developments to associate to each
bidimensional vector u0 (eqn. 11.113) a regular lattice of points in the u, w plane
identified by the set of position vectors, forming a 'reciprocal lattice'
= u0 4- ptx + qt2 (11.134)
with p and q arbitrary integers and t\ and t2 defined in eqn. 11.132. For elements
in the equilateral triangular arrangement of Fig. 11.1, the coordinates of the points
(eqn. 11.134) are given explicitly by:

Uopq =
R*~fP* W
OPQ = (2^-p)~ (11.135)

To proceed, the following expansion of a 2-dimensional delta function (a general-


isation of eqn. 11.95) is needed:

47r 2 m t--oo n r.oo

x exp [/(H —uo)(msi + ns2)] (11.136)


where C is the area of the elementary cell of the element grid.
Insertingeqn. 11.136 in eqn. 11.131,
+
4n2 ~ VT , u

(11.137)
which is the result sought.
For the numerical evaluation of eqn. 11.37 a truncation radius in the u, w plane
264 Conformal arrays

is chosen. The summation will be extended to the points (eqn. 11.34) internal to
the circular region so defined. The truncation radius has to be chosen large enough
to ensure that, outside the circle, the modulus of the transform of each of the M
selected mode functions | efe(w) | is negligible with respect to its peak (say, — 20 dB).

11.4.4 Array RGP for an arbitrary excitation. Single excited element RGP
The evaluation of the RGP via the eigenexcitation method does not require the
inversion of large matrices. Also the RGP of the single excited element is the same
for every array element, neglecting edge effects. The computational procedure is
straightforward. The derivation of the expressions of the d and (j> components of the
RGPs is fairly complicated, however, and will be reported only in its general lines.
Deriving an expression for the RGP requires performing the following steps:
(a) Establish an expression for the tangential electric field on the cylinder pertain-
ing to each eigenexcitation.
(b) Determine the associated field for large p.
(c) Use superposition according to eqn. 11.121.

Step (a): The aperture distribution on the generic element 'm, «' for the eigen-
excitation e(u0) is clearly the following:

I [6oh +
Z1TL fess0

x exp [-/tf O '(/?»i + ns2)] (11.138)


The FT of the transverse electric field distribution for the eigenexcitation e(u0) is
from eqn. 11.138

x exp [-/(a -UoYmst + ns2)] (11.139)

Step (b): From eqns. 11.89, 11.90 and the periodicity of the structure, it follows
that the far field for eigenexcitation e(u0) is given by
vl/2
/ h
g(i,wo;r,6,0) =

V exp [~jr{wOpq cos d 4- sin d(k2 - wg pg ) 1/2 )]


|u; O p a !<fe {K ~

x
[^0 (wopa \ 0) + 4>S<f){uQPq; ^)] exp | — /i ($ — — j | (11.140)

where (see eqns. 11.134 and 11.135)


Conformal arrays 265

M-X
XI [5Os+rs(«o)]6*(«op«) (11.141)
s=0

-jexp\-jpL (0-^) M-i


(«o;/0) = w(2),
L \ 2 I [§o, + rs(«0)]
1

4- l*Pf\m esz(uopq)) (11.142)


/C WOpq J
(c): The evaluation of eqn. 11.121 requires an integration with respect to w 0 ,
that is done asymptotically for r -* °° via the saddle-point method. The procedure is
standard and the result is:

[ 2n( n\ |
/ / Q-mn sxp \jiL — I m lH— I + ; « M cos ^ (11.143)
m n [ \ ) J
The RGP of the single excited element in the reference position is obtained by
having in eqn. 11.43 all amn = 0 except fortfoo= 1.
In connection with the evaluation of eqn. 11.143 notice that the second sum-
mation is extended only to the wavenumbers in the z direction corresponding to
propagating harmonics.
11.4.5 Aperture matching and element pattern design
We want to show that the element RGP can be modified via an appropriate design
of the element matching network. A technique will be introduced, and its effective-
ness exemplified by several illustrative cases. Although more complicated schemes
can be developed, only the simplest case of matching networks identical for all
elements will be here discussed.
Consider the plane tangent to the array surface through the centre of the
reference element and, on that plane, the direction of the cylinder axis and the
circumferential direction, perpendicular to the axial one. Suppose that one seeks
to increase the element RGP in a direction forming the cosines 7 and a with the
axial and circumferential directions, respectively. To accomplish this purpose a
266 Conforma/ arrays

matching network, the same for each element, will be selected to match the array
when excited by the eigenexcitation identified by

WQ = ky- i = hoc (11.144)


R

Fig. 11.15 Matching and element pattern shaping network


n,_ lG0

Go = Gii, w0)
Bo = B(i,w0)

The element RGP will be matched also in the directions — y, — a, symmetric with
respect to broadside. The approach and the ensuing procedure is easily justified by
observing that the realised gain pattern of an element in an array on a cylinder
having a large radius is obviously 'similar' to that of the same element in a planar
array with the same element grid. For the latter case, the element pattern is maxi-
mised in a direction a, y if the array is made transparent, i.e. matched, for a plane
wave incident from such a direction, or, equivalently, if it is matched for the array
beam steered in the direction a, y. The extension of the procedure to a cylindrical
array will therefore consist of matching the array to an eigenexcitation have the
same relative phase progression among the elements. All our considerations will
apply to a narrow band, ideally monochromatic case.
To introduce the technique consider the terminal admittance for the funda-
mental excited mode pertaining to the eigenexcitation identified by eqn. 11.113.
This 'active' admittance is, of course, the result of the mutual couplings among
the array elements for the particular eigenexcitation considered. Its expression is,
clearly,

= Yo l l o ( " ° \ (11.145)
111 (w J
Suppose now one introduces in each element a matching network, consisting of
shunt susceptance — Bo located at half transmission-line wavelength from the
aperture — a waveguide iris — and a perfect transformer having a turns ratio equal
to n2:n1 = (Go^o) 172 (Fig. 11.15). Clearly this network matches the array for the
eigenexcitation e(u0). Assuming that higher-mode interaction with the matching
network is negligible, a reasonable assumption, the set of aperture voltages for the
various modes (eqn. 11.120) are multiplied by the factor y/G0/Y0.
Conformal arrays 267
Following this procedure the gain can be enhanced in prescribed directions off-
broadside, at the expense, of course, of broadside gain. The total effect is a flatten-
ing of the element pattern with respect to the case of array matched for equiphase
excitation (7 = w0 = 0).

Fig. 11.16 Aperture geometry and element grid for the numerical examples
Lattice 1 Lattice 2
h = 0.3\0 h = 0.28867 \ 0
d=1.039\0 d=1\0
a = 0.22\ 0 a = 0.22\ 0

11.4.6 Numerical examples and discussion


The eigenexcitation method of analysis and the effectiveness of the technique of
element pattern design will be illustrated by several numerical examples. Three
cylinder radii and two different lattices have been considered. The element aperture
radius for all cases has been assumed equal to 0.22 wavelengths at centre frequency,
(Fig. 11.16). The element waveguides are assumed to be filled by a dielectric with
e = 2.5 (for e = 1 the fundamental modes would have been below cutoff).
A first set of figures refer to arrays with the lattice 1 of Fig. 11.16. A planar
array with the same lattice would have grating lobes in the visible space for certain
scan conditions. The first array studied has a radius of approximately 100 wave-
lengths. Circumferential and axial element realised gain patterns are shown for
nominal axial and circumferential polarisations, the elements being matched for
the equiphase conditions, conventionally called 'broadside' (Figs. 11.17 and 11.18).
For this lattice size and orientation a notch is present for the circumferential
polarisation in a direction approximately corresponding for a planar array to a scan
direction for which a grating lobe enters into visible space (Fig. 11.17). Characteris-
tically, however, the notch is absent for axial polarisation (Fig. 11.18). Also, the
slope of the realised gain pattern in the shadow region is substantially greater for
axial polarisation. In the axial direction the element realised gain pattern is very
insensitive to the cylinder radius and practically coincident with that for a planar
array.
268 Conforms/ arrays

By using a smaller size lattice (lattice 2 of Fig. 11.14) the patterns do not have
resonance notches but only a slight dip before 90°. For an array of approximately
50 wavelengths diameter, Figs. 11.19 and 11.20 show the patterns and also
illustrate the effect of 'off-broadside' match. For circumferential polarisation the
effect on the element RGP of a match for i/R = 2n sin 80°/X and w0 = 0 is an
increase of the gain off-broadside. At 80° the gain reduction with respect to broad-
side is approximately 6dB, instead of the 12dB for the more conventional equi-
phase match. For the RGP of an axially polarised element the match off-broadside
produces similar effects (Fig. 11.20).

100 J
30 60 90 120 150 180
4> t degrees

Fig. 11.17 Element realised gain pattern: Circumferential plane; circumferential polaris-
ation; equi-phase match (R = 99.23X)

In Fig. 11.21 the realised gain patterns in the circumferential plane for three dif-
ferent match conditions are shown for a cylinder of 1IX radius; the pattern for
frequency variations of ± 5% with respect to centre frequency are shown in Figs.
11.22 and 11.23. Broadside and 80° off-broadside match conditions have again
been considered. A notch begins to appear at high frequency. The modification of
the realised gain pattern with frequency depends substantially upon match con-
ditions, and thus is not due simply to geometrical reasons. In calculating the curves
of Figs. 11.22 and 11.23, the matching susceptance shunting the aperture has been
assumed constant with frequency (a reasonable approximation in a ± 5% band).
The element RGP in a conformal array is only an indirect indication of the scan
coverage capability (unlike the planar case). The influence of the array geometry,
Con formal arrays 269
lattice and match condition is illustrated in Fig. 11.24 to 11.28. Two arrays have
been considered, having 37 and 313 elements, respectively. The cylinder radius is
equal to approximately 11 wavelengths and the element lattice is the one denoted

30 60 90 120 150
^ , degrees
Fig. 11.18 Element realised gain pattern: Circumferential plane; axial polarisation; equiphase
match {R = 99.23X) (Lattice 1).

by '2' in Fig. 11.16. The projected apertures of the arrays, in the direction of the
normal to the plane tangent to the array at its centre, are circular. The projected
illuminations are Gaussian with a lOdB taper. The azimuthal sectors occupied by
the arrays on the cylinder are equal to 25° and 56° for the small and the large
array, respectively. The array pattern is calculated by using expression 11.143. The
free array excitation is chosen always in such a way to have the radiative contribu-
tions of all elements adding in phase at the peak of the beam. Since the phase
patterns of the element realised gain are essentially flat up to 80° from broadside,
the phases of the excitations turn out to be practically coincident with those
obtained from simple geometrical optics considerations. In Fig. 11.24 the realised
gain pattern of the array with 313 elements is shown, for broadside scan. The maxi-
mum gain is approximately 30 dB for equiphase match. Two different matching
conditions have been considered: equiphase and i/R = 2n sin 80°/X (with w0 = 0).
The two patterns are very similar, as expected, the differences lying in different
peak gains with the far-out sidelobes slightly higher for the case of match off-
broadside, because of the less efficient filtering effect of the element pattern.
270 Con forma I arrays

Figs. 11.25 and 11.26 show the envelopes of the maximum array gain vs. scan
angle for linear polarisation in the plane of scan and different match condition for
the array with 37 elements. Figs. 11.27 and 11.28 show similar curves for the large
array on the same cylinder. For match off-broadside the gain drop-off at 80° is
approximately 6 dB for the circumferential plane and 7 dB for the axial plane. The
absolute gain at 80° is 2 dB greater than that which would be obtained with the
same elements broadside matched.

30 60 90 120 150 180


0 , degrees

Fig. 11.19 Element realised gain pattern: Circumferential plane; circumferential polarisation
(R = 49.33X) {Lattice 2)
match at HR = 2*7X sin 80°
equiphase match

11.5 GTD evaluation of mutual admittances: Generalisation to a generic convex


surface

The methodology developed in Section 11.3.4 for the calculation of the generalised
mutual admittances can be formally extended to arbitrary convex surfaces. The
following GTD formulation is based on the work of Lee et al.15
A point of the surface is again identified by a position vector s. The large
parameter in the asymptotic expression of the Green's function is:
m(s) = tikRt(s)}1/3 (11.146)
Con formaI arrays 271
in the cylindrical case this parameter being a constant along a helical path. In eqn.
11.146, Rt(s) is the radius of curvature at the point s in the direction of the surface
ray; that is in the direction of the tangent t. The radius of curvature in the direction

CD

10

•I"

5 30
O

30 60 90
TT/ 2 - © , deg

Fig. 11.20 Element realised gain pattern: Axial plane; axial polarisation (R = 49.33X)
(Lattice 2).
match for w = 2n/X sin 80°
equiphase match

match at ///? = — sin 70°


ID

c 10-

g. 20
c
'§• 30 match at i/R = —sin 80°
•o
w 40
. equiphase match

* 50
0 30 60 90 120 150 180
$ t degrees

Fig. 11.21 Element realised gain pattern: Circumferential plane; circumferential polarisation
(R= 11.6U)

of the binormal b will be denoted by Rb(s). By extending the definition (eqn.


11.57) a distance parameter is introduced:
1/3

ds (11.147)
272 Conformal arrays

the integral being extended to the ray geodetic path. Another parameter of interest,
entering in the expression of the surface magnetic field, is
1/2
ks
(11.148)
2m(s')m(s")T

m \

10
E
to

Q. 20 \
c
d
01
30
~o
to
V
in
0 40 \\

\
f ' '

(3 30 60 90 120 150
<f> t degrees

Fig. 11.22 Element realised gain pattern. Frequency sensitivity {Lattice 2)


Circumferential polarisation
Circumferential plane
Equiphase match, R = 11.61X
_ . _fo
fo-0.05fo
fo + 0.05fo

depending upon the curvature at the source and observation point. The parameter
T\ is unity for the special case of constant ray curvature. Finally, the mean radii
of curvatures will be introduced:
RtV,*") = [Rt(s')Rt(s")]1/2 (11.149)
Rb(s',s") = [Rb(s')Rb(s")]l/2 (11.150)
f
Denote by s(s',s") the shortest distance between the points s and s" along a geo-
detic path. Redefine for the general convex surface the functions Ht(s) and Hb(s)
as follows:

Hb[s(s',s")] = G(s) ( l - f A
(11.151)
Conformal arrays 273

Ht[s{s',s")} =

(11.152)

90 120
t degrees

Fig. 11.23 Element realised gain pattern. Frequency sensitivity (Lattice 2)


Circumferential polarisation
Circumferential plane
Match at /"//? = 27r/X sin 80°'
fo
fo-0.05fo
fo + 0.05 f o

with G(s) given by eqn. 11.58. When the surface is cylindrical eqns. 11.151 and
11.152 reduce to the appropriate expressions given in Section 11.3.4.
Consider a small pencil of surface rays originating ats' and propagating towards
s". The divergence factor/^ of the pencil is defined by (Fig. 11.29):
sdxjj'
D(s,s ) = (11.153)

where pc is the caustic distance of the wavefront at s" and is always positive.
Denote by i (s',s") the unit vector tangent to the ray path at s' from a generic
point of aperture '/?' to a generic points" of aperture 'q\ In a parallel way i (s",s')
is defined. Expression 11.67 for the mutual admittance between two apertures
for the ith and kth modes is generalised for an arbitrary convex surface as follows:

d£dV'i(s'\sf)Hb[s(st,sff)]D(s\sff)

x [?(s\s") V(s')] + JJ d$"dri"ek(s") • j"J d%'dn'b(s",s')Ht[s(s,s")]

xD(s',s")[b(s',s"ye\s')] (11.154)
274 Con formal arrays

where p and q identify the two apertures considered and %', rf and if, 17" are two
planar rectangular systems on the planes tangent to the surface at the centres of the
aperture *p' and 'q\ respectively.

£D
T3

- 50.00

-60.00
0.00 20.00 40.00 60.00 80.00 100.00 120.00
<p t degrees

Fig. 11.24 Array realised pattern. Circumferential plane; circumferential polarisation.


Number of array elements = 313; R - 11.61\ {Lattice 2)
Equiphase match

Match at — = — sin 80°


R X

11.6 Conical arrays

11.6.1 Self and mu tual admittances


The formulation discussed in the previous Sections can be directly applied to arrays
of apertures on a cone> With reference to Fig. 11.30 let ru 0 t and r 2 , <j>2 be the
distances from the cone tip and the azimuthal angular coordinates of two generic
points s\ s" of the first and the second aperture, respectively. If 60 is the semi-
aperture of the cone, the distance s between the two points along the shortest geo-
detic path is
s(s\s") = \/rl + r\ - 2rtr2 cos (A sin 0O) (11.155)
where A is the angular separation between the two points assumed positive if s" is
reached from sf moving counterclockwise. The vector i (s\s") tangent at s' to the
ray path from s to s" has components in the direction (j> and — r
(11.156)
where
Conformal arrays 275

(11.157)
s(s\s")
with — 7r < r2(s', s") < 7T. Similarly,

0 10 20 30 40 50 60 70 80 90
4> t degrees

Fig. 11.25 Array scan coverage. Circumferential plane; circumferential polarisation match for
HR = k sin 80°. Number of array elements = 37; R = 11.61 \ (Lattice 2)
/ 2n
Match for — = — sin 80
R \
- Equiphase match

t(s\s) = [sin n(s",s'), cos Sl(s\s')] (11.158)


where
. , n i, rx sin (A sin d0)
sin 12(5 , s ) =
o
(11.159)

If, at each point of the geodetic path, i is expressed via its orthogonal components
along two orthogonal axes on the tangent plane in the circumferential direction 0
and in the direction — f (towards the cone tip)
(11.160)
then
b = (f r ,*0) (11.161)
Also, see eqns. 11.149 and 11.150,
\fr\r~i tan 60
R+ = (11.162)
I sin £2(s', s") sin £l(s", s')\
276 Conformal arrays

tan 60
Rh = (11.163)
cos Q(s , s") cos O(s", sr)\
f
Also
T(S,S") = [hkrl\smQ,{s,sft)\smd0]l/3\A\cos2/3e0 (11.164)

0 10 20 30 AO 50 60 70 80 90
— - 9 , degrees from broadside

Fig. 11.26 Array scan coverage. Axial plane; axial plane match for w = k sin 80°. Number of
array elements = 37; R = 11.61X (Lattice 2)
Equiphase match
Axial plane match

Tl(s',s") = (kslT)v\2k2rir2yV6 [sin n,(s',s") sin £l(s",s') cos0 o ] 1/3 -


(11.165)
On the cone
Df = 1 (11.166)
From eqns. 11.151 and 11.152 for ks > 1 and n ( / , s " ) , £2(s",s') not too close to
zero or it, by using the residue series for the Fock function (see Appendix 11.8.3
and eqns. 11.65 and 11.66.)
l sin n(5 f ,s^) sin 12(s",s')| cot 6O]1/3

x exp - 0 . 8 8 r - / [j- 4- 0.5lr 4- ks (11.167)

(11.168)
In applying eqn. 11.154, only the shortest geodetic ray needs to be considered.
However, if | A| ~7r, two surface ray contributions should be taken into account.
Higher-order geodetic ray contributions are negligible for structures large in terms
of wavelengths, i.e. those for which the formulation applies.
Conformal arrays 277

To conclude this brief discussion on admittance calculations for arrays on


conical surface, certain final remarks are in order.
(a) The cone tip scattering has not been taken into account. In many practical
cases the maximum arc length separation among array elements is much smaller
than the distances of the elements from the cone tip. An approximate expression
for the contribution to the admittance due to the scattering of the cone tip is given
by Lee,13 and is based on the work of Primdore Brown, et al.11 and Primdore
Brown.18 For two equal rectangular slots of dimensions 2a and 2b, with centres at
distances rx and r2 from the tip and with their wide sides forming angles coi and
a?2? respectively, with the radial direction on the cone surface, the extra term in the
admittance expression for the TE10 mode is

i2= Fsin sin co2 (11.169)

0
2
4
6
8
10

12

14

16
10 20 30 40 50 60 70 80 90
f t degrees

Fig. 11.27 Array scan coverage. Circumferential plane; circumferential polarisation match for
HR = k sin 80°. Number of array elements = 313;/? = 11.61X {Lattice 2)
Circumferential polarisation match
Equiphase match

where T is

4(ab) /tan0 o \ (sin kb


T = o0(d0) (11.170)
sin0 o \ 2?r / \ kb
and the tip diffraction coefficient o0(do) has the approximate expression
o0(60) = A(do)exp[jB(0o)] (11.171)
where
A(d0) - 1.3O570Q 1 -1.755+ 2.7720Q-1.45902 (11.173)
278 Con formal arrays

B(0o) = 2.7195+ 1.46O80Q-1.129502+ O.65660-J (11.174)


(60 being in radians).
(b) The GTD formulation for the self admittances leads to a divergent integral due
to an illegal exchange of an integration and a differentiation.13 A procedure to
overcome this difficulty consists of considering the self admittance as a sum of the
self admittance for the aperture on a ground plane plus a correction term. An
alternative procedure consists of considering an equivalent cylinder having a radius
equal to the 'local' radius of the cone cross-section at the location of the element
considered. Then the self admittance for the cylindrical geometry is calculated via
the harmonic series as discussed at the end of Section 11.32.

^ - 0 , degrees from broadside

Fig. 11.28 Array scan coverage. Axial plane; axial polarisation match for w — ks\nS(f.
Number of array elements = 3 1 3 ; / ? = 11.61X (Lattice 2)
Axial polarisation match
— Equiphase match

11.6.2 Radk tion pattern


For the conical geometry also it is possible to establish formally a modal field
expansion matching the element-aperture magnetic currents. Although the
approach is relatively straightforward in principle, the numerical complexity
makes the method very impractical. In fact, few examples are reported in the
literature of a successful application of the approach to a realistically complex
case of pattern calculation (see, for instance, Balzano and Dowling2) except for
azimuthally uniform rings of magnetic current.17 The modal expansion for the
conical geometry will not be discussed here because it can be found in textbooks
(for example, Harrington10). It is only recalled here that computational prob-
lems exist, related to the very slow convergence of the modal series and also to
the need for determining the roots of the associate Legendre functions P™(cos 0O)
and their derivatives d / ^ c o s 6)1 dd equated to zero for 6 = 60.
Conformal arrays 279

The approximate method of pattern calculation presented below is based on an


adaptation to the conical geometry of the expressions established for the cylindrical
case and is derived from simple heuristic arguments. In the discussion that follows it
is assumed that the magnetic-current distribution is a known quantity having been
determined by using the methods of the previous Sections. Again, lit region and
penumbra and shadow region are considered separately.

Fig. 11.29 Divergence factor for a pencil of rays originating at s and propagating towards s"

(a) Lit region: As for the cylindrical case the smallest azimuthal separation
between the aperture angular location and the direction of observation is ± |i^ol>
with the positive sign if the direction of observation is reached from the aperture
moving counterclockwise along the shortest possible arc. A rectangular system of
coordinates has the origin at the cone tip, with the z axis coincident with the cone
axis. In this reference the vector sp identifying the location of thepth aperture has
the expression

sp = (11.175)
Because of the structure rotational symmetry only the difference between the azi-
muthal angular coordinate of the direction of observation and the azimuthal
angular coordinate of the aperture appears in the radiation pattern. Thus one can
assume that in eqn. 11.175 yp — 0, that is, consequently, <j> = ^ 0 . Then the propa-
gation vector k in the direction of observation has the following components:
kx = k sin 6 cos i//0 (11.176)
ky = k sin 0 sin i//0 (11.177)
280 Conform a I arrays

kz = k cos 6 (11.178)
Let J, 7? be two orthogonal coordinates on the plane tangent to the cone at the
aperture centre, (Fig. 11.31). The unit vectors | , T) in the direction of the | and 17
axes point in the circumferential direction and toward the tip of the cone, respect-
ively. The unit vectors x, y, z are related to | , 7? and to the unit vector h in the
direction of the normal to the aperture by the transformation
i = y (11.179)
7? = — x sin 60 4- z cos 60 (11.180)
h = x cos 60 -\- z sin 0O (11.181)

Fig. 11.30 Geometry on the developed cone

Extending the result established for the cylinder in the lit region, the radiation
pattern is approximated by that of an aperture in a plane tangent to the cone at
the element location, and is given by
e~jkr
p
r->oo' r

where tp(u, w) is the FT of the aperture transverse electric field distribution with
respect to £, 17 (see eqn. 11.88). Explicitly the arguments of e p in eqn. 11.182 are:
fc'l = k sin 0O sin \jj0 (11.183)
Con formal arrays 281

= — k sin 9 sin 0O cos i//0 + k cos 0 cos 0O (11.184)


To determine the 6 and 0 components, consider the expressions of the unit vectors
in 9 and 0 directions
9 = cos 0 cos i//0Jc + cos 0 sin i//ojp — sin 0z (11.185)
0 = — sin \jjox 4- cos (11.186)

direction of
observation

geodetic paths
to diffraction
points

Fig. 11.31 Conical geometry. Radiation mechanism

The 0 component of eqn. 11.182 by using simple vector identities is found to be

6)-[«i>(*-i*-T?)x/2] (11.187)

or
(11.188)

that is, by using eqns. 11.179-11.181 and 11.186


282 Conformal arrays

e~Jkr
Eg(r, 6,0) ~ - / * e1k-*P(- fj cos \p0 + | sin 60 sin i//0) *P(k• i * '*?)

(11.189)
or finally
e-jkr

x ["" enp(k* 5» **^) c o s ^o + %>(£•§, fc-r?) sin ^ 0 sin 0O] (11.190)

In a completely parallel way the 0 component is found to be


ke~jkr t f
$(6 0 ) / ^ ' [ ( ^ i

cos ^o) + € nP(km%,kmfj) cos0 sin \|/O] (11.191)


For a half-cone, angle 60 = 0, eqns. 11.190 and 11.191 reduce to the expressions
11.99 and 11.100 established for the cylindrical geometry, as expected.

(b) Shadow and penumbra region: The electromagnetic field radiated from an
aperture propagates along geodetic rays on the surface with a decay due to the loss
of energy continuously radiated in the direction of the ray tangent (Section
11.3.5). Thus to each direction of observation certain geodetic paths are associated,
having at certain 'diffraction points' their tangents pointing towards the direction
of observation. As for the cylindrical case two geodetic rays only will be assumed to
contribute significantly to the pattern. They are associated with the azimuthal
separation \jj+ and $' = — |^~| between the aperture and the direction of obser-
vation, the positive angle corresponding to a motion in counterclockwise sense. The
angular separations between the aperture centre and the two diffraction points are:

A+ = ¥~~: (11.192)
and
A" = i//~-f| (11.193)

The angle y formed by the direction of observation and the generator of the cone,
the same at both the diffraction points, is given by (Fig. 11.31)
cos 7 = cos 6 cos 0O (11.194)
The directions of the geodetic rays from the aperture centre to the diffraction
points are identified by the angles £2+ and O" between the geodetic rays and the
direction of the cone generator. Their expressions are
£1+ = 7 - A + s i n 0O (11.195)
£T = - 7 - A " sin 0O (11.196)
Conforms/ arrays 283

The distances of the two diffraction points from the tip of the cone are found to be

4 = (SUIT)" 1 sin(l7=FA ± sin0 o l)'p (11.197)


Assume again that the centre of the pth aperture is at yp = 0, see subsection (a).
The diffraction points are identified by the position vectors sj, sj, whose expressions
in the rectangular reference are

Sd = /*J[sin(9o(iccosA ± +>'sinA ± )]+zcos^o (11.198)


The geodetic distances between the aperture centre and the diffraction points are
given by:
s(sp,s$) = yjrl + {4? - 2rprl cos (A 1 sin 0O) (11.199)
Denote by fr the unit vectors in the direction of the normal to the cone at the dif-
fraction point. In the JC, y,z reference fr are represented as follows:
rf = ±cos0 o (xsin0-->>cos0) + sin0oi (11.200)
having assumed the aperture at yp = 0. The unit vector tangent to the geodetic ray
at the diffraction points has the expression

i = sin 0 cos 0jc + sin 0 sin <j>y 4- cos Bz (11.201)


where 0 and 0 identify the direction of observation. The binomials at the diffrac-
tion points are given by
b± = i x/r (11.202)
that is
o = (± cos 0 cos 0O cos 0 4- sin 0 sin 0O sin 0) x + (— sin 0 sin 0O cos 0
± cos0 cos 0o sin0)j> +sin0 cos0 o f (11.203)
In order to generalise to the present geometry the expressions 11.101 and 11.102
established for the cylinder, by using eqn. 11.164, the length parameter is introduced
r* = 7(^,5*) = [|A:/- p |sin^ ± |sin0 o ] 1 / 3 |A ± |cos 2 / 3 0 o (11.204)
Also, the radius of curvature will be replaced by its average value. By using eqns.
11.162 and 11.197, the following substitution is made:
1 kR 1
i,n± foptanflp

It is convenient to consider the contributions of the two rays from each diffraction
point separately, and using eventually superposition. For each contribution the
components in the directions of the normal and binormal to the geodetic rays,
given by eqns. 11.200 and 11.203, will be considered. It will be apparent that this
reference is the most 'natural' to express the field. The 0 and0 components of the
284 Conformal arrays

pattern will then be obtained via a coordinate transformation. The component in


the direction of the normal to the diffraction point is obtained by a generalisation
ofeqn. 11.101 as follows:

E*±(r, 0, 0) ^ / — exp {-f[kr + s(sp, s*)] + fk-s&}

± j(hkRj)1/3 cos ^ V 1 ) ( r ± ) € T?p(^ sin £2*, k cos Or)}, (11.206)


The component in the direction of the binormal is, from eqn. 11.102,

E^(rJP)^}'xp{

x sin Qr(rpr±dyll2eriP{k sin £2*, k cos n * ) / ^ * ) . (11.207)


The 0 and 0 components are then obtained via the following coordinate trans-
formation:
Eg± = (^•/z ± )^± + ( 0 - ^ ± ) ^ ± (11.208)
± ±
E%± = (4>-h )E%± + (])-b )EZ± (11.209)
where
Q'fr = - sin 0 sin 0O (11.210)
O-b* = ±cos0 o (11.211)
0-w* = +cos0o (11.212)
0-6* = - s i n 0 s i n 0 o (11.213)

11.7 Acknowledgment

The GTD formulation for the surface magnetic field, discussed in Sections 11.3.3
and 11.3.5 has been contributed by S. W. Lee. Sections 11.3.4 and 11.6.1 also are
based in part on his material.

11.8 Appendices

11.8.1 Modal admittances for rectangular and circular waveguides


Consider a waveguide of arbitrary cross-section. Denote by $ the propagation
constant for a certain mode. If fc is the cutoff frequency, @ is given by:
Conforma/ arrays 285

„-.
Lw
f>U (11 .214)

f<U (11 .215)

where
k = coy/ql (11.216)
is the wavenumber in an unbounded medium having the same permittivity e and
permeability JLI of the material filling the waveguide. The modal admittances are
then given by the following expression for the TE and TM modes, respectively:
TE modes:

rTE = M_ /I (11.217)
k
TM modes:

yTM - !L I?L (11.218)

For a rectangular waveguide with sides having the lengths 2a and 2b, the cutoff
frequencies are given by the following expressions, valid for both TE m n and TM mn
modes:

(11.219)

Consider a circular waveguide having a cross-section with the radius 'a\ In the
following, as customary, the first of the two indexes 'identifying a particular mode,
refers to the azimuthal periodicity. The cutoff frequencies are given for TE modes
by:
(/ c )™ = -$*= (11.220)
2iray/€fJ.

where x'np is the pth root (ordered with increasing magnitude) of the equation:
dJn(x) = 0 (11.221)
dx
Jn being the Bessel function of the first kind and order n. For TM modes, the cut-
off frequency is given by:

(f\TM = (11.222)

where xnp is the pth root of the equation:


Jn(x) = 0 (11.223)
286 Conformal arrays

11.8.2 Fourier transforms of the vector mode functions for rectangular and circular
waveguide radiators
Consider the 'reference' element whose centre is located at £ —z = 0 (Fig. 11.6).
For a rectangular aperture with % and z dimensions equal to 2a and 2b, respectively,
the vector mode functions can be written as follows. For TE nm modes,
,2-1-1/2, x l/2

(n.224)

where rn is the Newman factor equal to unity for n — 0 and equal to 2 otherwise,
and
[ t \ [muz mir\
( ) (11-225)

(11.226)

Similarly, the TMnm vector mode function is:


,1/2
\x
mir 'm
2b I
/ J \ /
WITT I
(11.227)

From eqns. 11.224 and 11.227 it is apparent that in order to establish the expression
for the FT of the vector mode functions, the FT of eqns. 11.225 and 11.226 have
to be established. This is quite elementary, the results for the transform of eqn.
11.225 being:
1
nm, ^ \ ,
FZ (u, w) = — exp
P — }(m + n 4- 1) —
2* [ >2
Inn

mix

(11.228)
Conformal arrays 287

The expression for the FT F" m (w, w) of eqn. 11.226 is obtained from eqn. 11.228
by using the following replacements:

u <- w

Then from eqns. 11.224-11.228 the FT of the % and z components of the various
vector mode functions are immediately obtained.
Consider now a circular waveguide aperture element with the radius of the cross-
section denoted by a. Introduce a planar polar system of coordinates p, i// on the
aperture of the waveguide element, associated with the rectangular system | , z via
the standard transformation

p = 4- z 2 ; cos \jj = sin \p =

The radial and circumferential components of the orthonormalised vector mode


functions have the following expressions. For the TEni> modes (see Appendix
11.8.1 for the meaning of the symbols)

sin

(11.230)
and for the TMnp modes
^npp\ fcOSMi//)

01-232)
TI pxnpJn_x{xnp) \ a
In the plane of the wavenumbers M, W introduce a planar polar system t, /i with

t = COSM = — ; sin// = — (11.233)

It is convenient to introduce the radial and circumferential components of the FT


of the vector mode functions. As discussed in reference 5, they are obtained by
Fourier transforming linear combinations of eqns. 11.229 and 11.230 or 11.231
and 11.227, specifically:
288 Con formal arrays

j o J o [Ep(p, \j/) cos (ty-fi-Ey (p,\l/) sin ( ^ - /


x exp/[p/ cos (i// — fi)] pdpd \jj.
for the radial component, and

- . . . -u,,-, ^sin(i//--ju) + i}l/;(p, y)cos(i//-/

x exp/fp^ cos (i// — fj.) pd pd 4?.


for the circumferential component. The detailed evaluations are omitted here. The
results are as follows. For TE np modes:
JnnK
(at)} (cosnn)
' ' (11.234)

] (H.235)

For TMnp modes

(n.236)

e?{t,ii) = 0 (11.237)
the latter equality holding for all pairs of indexes n and p.
The | and z components of the FTs are obtained from eqns. 11.234-11.237 via
the coordinate rotation
M (11.238)
e2np(w, w) = et(t9 fi) sin M + eM(f, ju) cos /x (11.239)
In the text the alternative notation e^ has been also used for e^.

11.8.3 Fock functions


Fock functions have been widely used in diffraction problems. Only the basic
notions needed for their numerical evaluation are recalled here. Their use in electro-
magnetic theory is discussed in the collection of the lucid original papers by Fock9
(see also Bowman et al.6).
Using Fock notation, introduce the functions
w2(t) = yfi[Bi(t)-J'Ai(i)] (11.240)
Con formal arrays 289

W
2(t) = — (11.241)

where Ai(t) and Bi(f) are the Airy functions whose definition and discussion is in
Abramovitz and Stegun1. The general definition of Fock functions is not given
here, but only those functions used in the text are here listed:

v(x) = _ L e / ( * , 4 V , 2 J r ?&e-»*<lt (11.242)

(11.243)

1 r e-Jxt
r \ d (11.245)
7T ' r W ( 0

gu\x) = —L f —,— dt (11.246)


\JTI J v vv2(r)
where the integration contour comes from — °° in the angular sector defined by
— 7r < arg(r) < TT/3. Notice that in Bowman et al.6 the Fock functions have been
defined as the conjugate of those above because of the different choice of the
time dependence exp (— /cor) instead of exp (/cor).
The numerical evaluation of eqns. 11.242 and 11.243 for positive values of x
is needed for the calculation of mutual admittances. The evaluation of eqns. 11.244
and 11.246 for small negative x and for positive x is related to the evaluation of
radiation patterns (Section 11.3.5).
For large positive value of the argument, the residue series representation is
rapidly convergent. If an and an are the roots of the equations
Ai(-an) = 0; Ai\-a'n) = 0 (11.247)
the residue series expressions for eqns. 11.242,11.243 and 11.244,11.245 are:

v(x) = yfte'H7r/4)x1/2 X (O""1 exp[-/xa^(exp-/7r/3)] (11.248)


n

u(x) = 2V^ i7r/4) x 3/2 £ exp[-/ra n (exp-/7r/3)] (11.249)


n

f(x) = eJW3) ^ [i4,-'(-aB)]-1exp[-/x«B(exp-/ir/3)] (11.250)


290 Conformal arrays

g(x) = <T>(7r/6)5 I [Ai(- a ^ K ] - 1 exp [ - / x o ^ e x p -/ir/3)] (11.251)

and the expression for # ( / ) is obtained by taking the derivatives of eqn. 11.251.
Numerical values of an and an are found in Abramovitz and Stegun,1 page 478.
For small value of the argument the functions (eqns. 11.242, 11.243) can be
evaluated by resorting to the following small argument expansion, reported by
Lee:13

*(*)- 1 -^-e'Wx™ +^ * 3 +^ e - ' W V ' J + ... (11.252)

1/2
+ n i 9STI
I 12 64
It has been indicated that the smoothest crossover between the residue series and
the small-argument expansions occurs for x ~ 0.6 (Chang et alP). The functions
/(*)> z(x) a n d g^'\x) are tabulated and graphs for g(x) and f(x) are in Bowman
et al.6, among others. Interpolation from the tables or from the graphs in the region
I x | < 1 (where the functions are smooth) is a procedure simpler than attempting a
Taylor expansion leading to rather involved expressions.
Tables of the Airy functions for moderate values of the argument are in
Abramovitz and Stegun.1
11.8.4 Evaluation of the radiation pattern of an aperture on a large conducting
cylinder

(a) Lit region: Denote by \p0 the quantity (fig. 11.12)

^o = 0-f? (11.254)
K
whose absolute value is less than n/2 in the lit region.
Consider first a positive \p0. From eqns. 11.97 and 11.98 the components of the
pattern of thepth aperture, located at %p, zp, are:
e-jkr l + oo r
££(r, 0, <t>) ~ exp (jkzp cos 0) exp \-jp UJ0 - ^ I

. P COS0 .. , ^., ,
-fccos0 + _ . 2 e z p \-,kcosS\\dv

(11.255)

Eg(r,d,(P)~-jy- J
r_>oo r nRsind -°° H^2)(kR sin 0)

X€ zp\^,k cos 0)dp (11.256)


Con formal arrays 291

where only the term n = 0 has been retained in 11.97 and 11.98. To perform an
asymptotic evaluation of eqns. 11.255 and 11.256 for large kR sin 6, the Hankel
functions and their derivatives will be replaced by their Deb ye asymptotic
expressions. The approximation is not valid in the immediate neighbourhood of
v~kR sin 0 (see for example, Sommerfeld19). It is argued, however, that the rapid
oscillatory behaviour of the phase of the integrand makes only a small region of the
v axis, where the phase is stationary, contribute significantly to the integral. This
region is far from v~kR sin 0. (The situation is, of course, totally different for an
observation point in the transition or shadow region.) On this basis the Hankel
functions in eqns. 11.255 and 11.256 and their derivatives will be replaced with
their approximations (Hasserjian and Ishimaru11)

HJ?\kR sin 0) ~ /-
' nkR sin 0 sin 0
x exp [-jkR sin 0(sin 0 - 0 cos 0)] (11.257)

Sin0)
" • ' — sin0
x exp [-jkR sin 0(sin 0 - 0 cos 0)] (11.258)
where
V
cos0 = yo n (11.259)
kR sin 0 '
Consider eqn. 11.256 first. By using eqn. 11.257 and changing the integration vari-
able from v to 0
e~ikr , ., ^ f /fffcR sin 0 sin 0
e

x exp jkR sin 6(sin 0 — 0 cos 0) — jkR sin 6 I \jj0 1 cos 0

x ezp(k sin 6 cos 0, k cos 6) kR sin 6 sin 0tf0 (11.260)


where the integration path in the 0 plane goes from — TT/2—/°° to TT/2 + /<». The
integral in eqn. 11.260 is now evaluated asymptotically for large kR sin 6, via the
saddle-point method. The procedure is standard and the result is

exp l — jkr + fk \zp cos 6 — R sin 6 sin j \p0


l 2
d<P)~jk ' '

x sin j\//o \ezp\ k sin 6 cos (i//o~T|>^ cos0| (11.261)

from which expression 11.99 is obtained.


292 Conformal arrays

If \po is negative, then

exp \-jv L-|Jl = exp \jv |ltfol + x II 01-262)

The integral in eqn. 11.256 becomes

ex ULl+^1 ex -'L\-?~
exp 7y +o exp
+oo [ \ ° 2/J /v \ f ° \ ° 2;
C S
J-~ H^2)(kR sin 6) 6zp\Ry ° / ^ J-~ H?\kR sin Q)
^ . f c c o s f \l ) ^ (11.263)
R
I
where the relationship
e*v*Hil\z) = /# 2) (z) (11.264)
has been used. The procedure for the derivation of eqn. 11.261 is repeated for this
case and leads again to the expression 11.99.
For the <f> component eqn. 11.255 by using eqn. 11.258, one obtains
e~jkr 1 f f [
%r, 0, 0) — exp ( jkzp cos 0) exp fkR sin 0 sin j3
r * Jc | |
sin 0

cos |3 cos 0 1
sin 0 cos 0, — k cos 0) H — — ezp(k sin 0 cos ]3, /: cos 0)
sin0
sin 0 1

fcsin0sin/3d/3 (11.265)
By applying again saddle-point integration
ke~*kr
%r, 0, 0) ~ e-JOr/2) s i n 0 e x p ( / f a ^ c o s fl)

cos
x exp —fkR sin 0 sin I \p0 — l \\e<pp \ksind
" f— I ^Ao — — i, k cos 0
/ ^X
4- cos h//0 cot 0e zp I A: sin 0 cos (i//0 - ^ | , k cos 01 [ (11.266)

which is the result quoted in eqn. 11.100.


(b) Transition and shadow region: It will be postulated that in eqns. 11.97 and
11.98 only the two terms for which the angle

\l/n = 0—£—2irn (11.267)


Con formal arrays 293

has an absolute value smaller than 2ir


Wn\<2n (11.268)
need to be retained in the summations, since it is argued that the values of the
other integrals are comparatively negligible because of the rapid oscillatory
behaviours of the integrands. The positive and negative angles i//n satisfying eqn.
11.268 will be denoted by \p+ and i//_, respectively. It is evident that
i//_ = I// + -2TT (11.269)
and
l i M + \p+ = 2TT (11.270)
Thus, the approximate expressions of the components of the radiation pattern
become:

e ! ( r+ ~exp
, 0, 0) ~ exp [ jkzp cos 6} f _
r nR
\J-°° H^2)\kR sind)
v
x |e0p

-exP
+ — H™\kR sin 0) r0pl/T*COS7 +
fcRsin20

xe* p k;,fccos0|| dv (11.271)


and

)--/-
vRsind [ J-°° ^ 2 ) ( / c ^ sin

(11.272)
In this case the asymptotically important contributions to the integrals in eqns.
11.271 and 11.272 come from the region of the v axis where v ~kR sin 6. In that
neighbourhood the Hankel function and its derivative can be approximated by a
combination of Airy functions and their derivatives as follows. Introduce the large
parameter
m(6) = (\kR sin 6)1/3 (11.273)
and make a change of variables

1
v-kR sin 6
= —Z7^ (11.274)
294 Conformal arrays

For small t, the following approximations hold (Bowman etal.6):

Hp\kR sin 0 ) - - $ ^ (11.275)


/nm(d)

sin 0 ) ^ ^ £ (11.276)
VW (0)
where the functions w2(t) and w ^ O have been defined in Appendix 11.8.3. By
using eqns. 11.275, 11.276 and 11.263 (with \\p_\ replacing i//0) eqns. 11.271 and
11.272 become

exp -/W* sin 0 L + - ^

exp

+ I1+ . ' )cot dezp(tm(6) + k sin 0, A: cos 0) \ dt

exp\-jkR sinJ \\;_\ *

: [e / r/w(0) , ^T \ / tm(0) \
</v>D k sin 0 , k cos d\— \ l -\
[ \ /? \ kR sin9J

(11.277)

with F a suitable integration path, see Appendix 11.8.3. Similarly fromeqn. 11.272,
exp (
e
/^ c °; ^M cJ-m sin e (r ^ |

w 2 (r)
X
^ P l - 1 1 1 ^ - A : sin (9,^ cos ( 9 ) ^ ) (11.278)
Conformal arrays 295

Since e^p{u, w) and ezp(u, w) are slowly varying functions of their arguments and
the main contribution to the integrals 11.277 and 11.278 comes from a neighbour-
hood of t - 0, no great error is committed if the argument of the FT of the aper-
ture field transverse field distribution is replaced by its value at t — 0. Also to form-
ally simplify the expressions, the Fock functions are introduced (Appendix 11.8.3).
By using the definitions, eqns. 11.258-11.260, the expressions 11.271 and 11.272
take the forms of eqns. 11.101 and 11.102.

11.8.5 Eigenexcitatiom of the periodic cylindrical array


It will be shown that eqn. 11.112 are eigenexcitations of the structure via a direct
verification of eqn. 11.115. From eqn. 11.112
L-i +°° rjj-
[Ykie(Uo)]mn = Z Z Y
m,p;n,q J ~ T exp - JU0' (psx + qs2)

(11.279)
Putting
m-p =s (11.280)
n-q = t (11.281)
eqn. 11.279 becomes, because of eqn. 11.110.

[Ykie(u0)]mn = Kl^j[™V--J'»o-(ms1+ns2)} £ £ tf?

(ssi + fs2)] (11.282)


which is the same as eqn. 11.115 if the identification is made
L - l +oo
Yki(u0)= I I tf'exp/«o•(», + »,) (11.283)

In the same way it is shown that the set of vectors e(w0) satisfy eqn. 11.116.
If the two wavenumber vectors u0, ux identify two eigenexcitations, to show
that the system eqn. 11.113 is orthonormal the following equality should be
verified:
h L-i +<*>
7TT Z Z exp-/(no-Hi)*(msi +ns2) =
27TL

(11.284)
where /, k and w0, w t are the azimuthal and longitudinal numbers for the eigen-
excitations. By recalling the expressions of the two periodic vectors, Si,s2, for the
equilateral triangular arrangement given by eqns. 11.108 and 11.284 becomes
296 Conformal arrays

+o L 1
° f } "
^ exp\f(i~k)jn-j(wo-W2)nh\ £
Y L J m=o

x e x p | / ( / - & ) — m\ = 6 l 7 l 5(w 0 -w 1 ) (11.285)

with the longitudinal wavenumbers w0, wt in the interval —n/2h<w<n/2h.


It is immediately verified that

exp / ( j - / c ) — J = Sik (11.286)

Therefore to validate eqn. 11.284 one needs to show that

— I exp[-f(wQ-wl)nk] = 5(wo™w1) (11.287)


2n n= _oo

in the interval ix/h < w 0 — Wi <7r//z. This equality, however, is proved by expand-
ing in a Fourier series of period 2TT//I the delta function in the right side of eqn.
11.286.
The formal statement of completeness for the system of vectors (eqn. 11.112) is
written as follows:
2TT ,

2
r exp [-jwo(n -p)h] dw0 = 8np8mQ (11.288)
J-irl2h

with m and q ranging from 0 to L. Through simple manipulation eqn. 11.288


becomes the identity

sin ( / i - p ) ^ L_!
J
I exp/i

and this shows that the eigenexcitations form a complete set and every array
excitation can be represented as a linear superposition of them.

11.9 References

1 ABRAMOVITZ, M., and STEGUN, I. A.: 'Handbook of mathematical functions'. Dover


Publications, 1965, Chap. 10
2 BALZANO, Q., and DOWLING, T.: 'Mutual coupling analysis of arrays of apertures on
corns', IEEE Trans., AP-22, 1974, pp. 92-97
Con formal arrays 297

3 BORGIOTTI, G. V.: 'A novel expression for the mutual admittance of planar radiating
elements', IEEE Trans., AP-16, 1968, (3). pp. 329-333.
4 BORGIOTTI, G. V., and BALZANO, Q.: 'Analysis and element pattern design of periodic
arrays of circular apertures on conducting cylinders', IEEE Trans., AP-20, 1972, pp.
547-555
5 BORGIOTTI, G. V.: 'Modal analysis of periodic planar arrays of apertures', Proc. IEEE,
56,1968,pp.1881-1892
6 BOWMAN, J. J., SENIOR, T. B., and USLENGHI, P. L. E. (Eds.): 'Electromagnetic and
acoustic scattering by simple shapes'. North-Holland, 1969
7 CHANG, Z. W., FELSEN, L. B., and HESSEL, A.: 'Surface ray methods for mutual coup-
ling in conformal arrays on cylinder and conical surfaces'. Polytechnic Institute of New
York, Limited Circulation Report, 1976
8 EDELBERG, S., and OLINER, A. A.: 'Mutual coupling effects in large antenna arrays',
IRE Trans., AP-8, 1960, pp. 286-297
9 FOCK, V. A.: 'Electromagnetic diffraction and propagation problems'. Pergamon Press,
1965
10 HARRINGTON, R. F.: 'Time-harmonic electromagnetic fields'. McGraw-Hill, 1961, Chap.
5, pp. 245-250
11 HASSERJIAN, G., and ISHIMARU, A.: 'Currents induced on the surface of a conducting
cylinder by a slot', /. Res. NBS, 66D, May-June 1962, pp. 335-365.
12 KUMMER, W. H.: 'Preface, special issue on conformal arrays', IEEE Trans., AP-22, 1974,
p.l
13 LEE, S. W.: 'GTD calculation of mutual admittance and element pattern of slot on con-
formal arrays', Dept. of Electrical Engineers, University of Illinois, Urbana, Illinois,
1979.
14 LEE, S. W., and SAFAVI-NAINI, S.: 'Approximate asymptotic solution of surface field
due to a magnetic dipole on a cylinder', IEEE Trans., AP-26, 1978, pp. 593-598
15 LEE, S. W., and MITTRA, R.: 'GTD solution of slot admittance on a cone or cylinder',
Proc. IEE, 126,1979, pp. 487-491
16 MONTGOMERY, C G., DICKE, R. H., and PURCELL, E. M.: 'Principles of microwave
circuits', Dover Publications, New York, 1965
17 PRIMDORE BROWN, D. C, and STEWART, G.: 'Radiation from slot antennas on a
cone', IEEE Trans., AP-20, 1972, pp. 36-39
18 PRIMDORE BROWN, D. C: 'Diffraction coefficients for a slot-excited conical antenna',
IEEE Trans., AP-20, 1972, pp. 40-49
19 SOMMERFELD, A.: 'Partial differential equations in physics', Lectures on Theoretical
Physics, Vol. 6, Chap. IV
20 'Special issue on phased arrays', Proc. IEEE, Nov. 1968, (11)
21 STEISKAL, H.: 'Analysis of circular waveguide arrays on cylinders', IEEE Trans., AP-25,
1977,(5), pp. 610-616
22 TILLMAN, J. D.: The theory and design of circular antenna arrays', University of
Tennessee Experiment Station, 1966
23 WATSON, G. N.: 'The diffraction of electric waves by the earth', Proc. R. Soc, A-95, pp.
83-99
Chapter 12

Circular arrays
D. E. N. Davies

12.1 Introduction

The field of circular (or cylindrical) arrays has received far less attention than that
devoted to linear and planar arrays. At first sight this seems surprising since one of
their principal advantages is the ability to deflect beams electronically through 360°
with little change of either beamwidth or sidelobe level. Circular arrays have been
applied, to a limited extent, in both communication systems and navigational aids
since the late 1930s.1 There has, however, been surprisingly little change in their
scale of application since that date, though there are now signs of a significant
increase of interest in such arrays for applications to DF (Direction Finding) and
ESM (Electronic Support Measures) systems.
The analysis and general understanding of the properties of circular arrays has
progressed steadily for many years and the most significant developments were
probably made by several groups of workers in the 1960s when the concept of
phase mode excitation was developed. The lack of applications in the past is prob-
ably due to the fact that the basic problems of exciting circular arrays with the
correct values of amplitude and phase are, in general, more complex than for linear
arrays. Furthermore, electronic scanning of directional patterns for circular arrays
may be difficult to implement and can require both the amplitude and phase of
each element of the array to be changed.
The basic symmetry of circular arrays offers a number of advantages, which in
addition to the 360° scan angle already mentioned can include an ability to
compensate for the effects of mutual coupling by breaking down the array exci-
tation into a series of symmetrical spatial components. This can also give rise to
directional patterns which remain constant in shape over broad bandwidths. These
advantages, together with new and more convenient methods of array phasing,
are beginning to increase interest in such arrays for a number of potential appli-
cations.
Early work on circular arrays related to their application for direction finding1
and their ability to produce omnidirectional coverage2 from a cylindrical con-
figuration of elements wrapped around a central mast or support structure. This
Circular arrays 299

latter configuration was applied in broadcast antennas to avoid shadowing from the
basic mast structure which supported the antenna. In this case controlled excitation
of the vertical aperture of the cylindrical aperture is used to achieve the desired
pattern in the vertical plane. This is generally achieved by stacking multiple rings
of circular arrays.
Other circular array applications have included Wullenweber arrays for direction
finding,3 wide bandwidth HF communication arrays,4 wrap-around antennas for
shipborne communications,5 navigational aids,6 spacecraft antennas,7 null steering
antennas for mobile communication applications8 and wide-bandwidth microwave
direction-finders.9
In the field of radar systems there have been several experimental circular arrays
aimed at achieving 360° electronic scanning10""14. These have been aimed at
evaluating different scanning concepts applied to fairly small apertures containing
perhaps 16 or 32 elements. The problems of high-power feeding and phasing for
circular arrays still represent a significant restriction on this type of application,
particularly for wideband operation. For the radar field there is, nevertheless, the
possibility of designing circular arrays for receiving only, where the problems of
phasing can more easily be achieved by array signal processing techniques.
There has also been a limited application of circular array configurations to
sonar systems15 although the published material available in this field is far more
limited.

12.2 Beam cophasal excitation

Probably the most obvious way of exciting a circular antenna array is to arrange the
signals from all the elements to add together in phase for a single (main beam)
direction. Assuming this direction to lie within the diameter plane of the array and
taking 6 = 0 as the main beam direction for the array shown in Fig. 12.1 requires
that the excitation at angle $ round the array of radius r is:

F(0) = exp(-/y-cos0j (12.1)

For omnidirectional elements and if the number of elements in the array is very
large (i.e. small interelement spacing), the normalised directional pattern in the
horizontal plane can be expressed as:
1 f27r \lnr 1
D(6) = — Jo F(0)exp; — cos (0-0)1 </0 (12.2)
giving

D(d) = J0\^-sin"-\ (12.3)


This directional pattern is shown in Fig. 12.2 for the case of a 5X diameter array.
It has high near-in sidelobes (— 9 dB) and a slightly narrower beamwidth than a
300 Circular arrays

corresponding linear array of length equal to the diameter of the circle. In practice,
the interelement spacing needs to be about A/2 or less in order to approximate
to this function.

I arbitrary reference axis

incoming wavefront

elements of receiving array

Fig. 12.1 Wavefront arriving at array from direction 6

in 9/2)

-10-

dB

-20-

-30
80 120
9, deg

Fig. 12.2 Directional pattern of phase-compensated circular array for a continuous aperture
distribution with no amplitude tapering and fora diameter of 5\

The reduced beamwidth and increased sidelobes are due to the increased con-
centration of elements at the sides of the circular aperture (compared with a linear
array). It has been shown by Fenby16 and by James17 that low sidelobes can only
be achieved by means of a symmetrical amplitude taper on the front half of the
circular array with very low or zero excitation on the back half of the array. These
results assume omnidirectional elements. Electronic beam rotation of such patterns
therefore requires electronic control of both amplitude and phase of each element
The vertical pattern (through the plane of the main beam) is also a J0(x) func-
tion of the form

D(a) = Jo | — ( 1 -cosa) (12.4)


A

where a is the vertical angle measured from the plane of the circle. Fig. 12.3 shows
such a vertical pattern for a 5X diameter array. Again this result assumes a con-
tinuous array.
The vertical beamwidth is rather large since this is analogous to the equivalent
Circular arrays 301

'end fire' position for a linear array. A single ring circular array therefore provides
a convenient-shape fan beam for rotation through 360°, although it is important
to realise that it is not possible to exercise independent control of the horizontal
and vertical beamwidths for a single ring array of omnidirectional elements. The
amplitude tapers designed to reduce the sidelobes in the horizontal plane will also
increase the beamwidth in the vertical plane.

J o [ 5 T T ( 1 - C O S CX)]

160

Fig. 12.3 Directional pattern in the vertical plane of a phase-compensated circular array of
diameter 5X

The most well-known use of circular arrays is probably for the Wullenweber
array which employs an arc of about 120° and excites it with beam cophasal
excitation.3 It therefore represents a phase compensated curved linear array. The
process of electronic beam rotation involves commutating the required aperture
distribution around the array to the appropriate grouping of elements. Wullenweber
arrays generally use mechanically rotating goniometers or corresponding electronic
switches. An organ pipe scanner could perform the same function for waveguide
feeds. Such arrays have been used for HF receiving antennas and for direction
finding.
The effect of directional elements on beam cophasal patterns has received only
limited study and the evidence suggests that directional elements would not be
particularly helpful, with the notable exception of the case of the restricted arc
excitation. Most Wullenweber arrays employ directional elements of a simple
cardioid or half-cosine shape. An advantage of operating untapered beam cophasal
arrays is that since all the elements have the same amplitude, electronic beam
rotation may be achieved by control of phase alone. However, the sidelobe levels
are too high for most applications. The rotation of tapered excitations usually
requires control of both amplitude and phase and this represents one of the limi-
tations of beam cophasal systems due to the cost and complexity of scanning.
A circular array of omnidirectional elements is very prone to mutual coupling
effects. On transmission it is even possible for one element to pick up more power
by mutual coupling than it is trying to radiate (i.e. negative radiation impedance).
The effect of mutual coupling in beam cophasal circular arrays can be minimised
by using directional elements and restricting the active arc.
302 Circular arrays

Recent work has also shown that by departing slightly from the beam cophasal
excitation for phase but avoiding any amplitude taper, it is possible to achieve side-
lobe levels below — 20 dB. This offers pattern rotation by phase variation alone.18

12.2.1 Wideband performance


If the phasing system of the circular array employs true time delays then the
direction and the/ 0 (X) shape of the main beam will be independent of frequency
and can be preserved over a wide bandwidth (until the interelement spacing
becomes too large, resulting in distortion of the pattern). Such arrays have been
constructed for the HF band and have worked over three octaves of frequency.4'19
The horizontal beamwidth of the array varies linearly with wavelength in the
normal way.
The main feature of beam cophasal patterns is that provided the element spacing
is small, in general less than A/2, the sidelobe pattern follows the/ 0 (x) (or tapered)
distribution over all angles in both dimensions. This means that there cannot be
any sidelobes higher than the first pair of sidelobes and the sidelobe levels will
decrease monotonically away from the main beam. The vertical pattern can there-
fore only be controlled by choice of the vertical directional pattern of each element
in the array, or by stacking circular arrays to form cylindrical arrays.

12.2.2 Electronic scanning for beam-cophasal patterns


A range of different phasing and switching schemes have been devised and tested
for electronic rotation of beam cophasal patterns for circular arrays. These include
diode matrix switches and mechanical rotating commutators and goniometers.3
The former approach leads to very complex switching techniques in order to rotate
the required amplitude and phase excitation, whereas the latter is restricted to
continuous scanning at rates set by the inertia of the mechanical parts. In most
cases, an arc less than 180° is excited with a suitable form of amplitude taper to
achieve low sidelobes. Several of these alternatives are discussed in Reference 20.
A rather different approach involves the use of a circular symmetric geodesic
lens such as a Luneberg lens.21 The output ports round the circular lens are con-
nected to elements of a circular array and one lens port is then used as a feed
element for the lens. Several types of lens can be used in this way as a feed medium
for a circular array.20'21 The above schemes provide a number of fixed beams from
the array and, in general, the performance approximates to a beam cophasal type
pattern due to sector excitation.
Fig. 12.4 shows the general schematic form for feeding a circular array from a
cylindrical geodesic lens. The pattern of the primary feed for the lens controls the
sector of the array to be excited and fixes the taper function.
Luneberg lens systems suffer from the fact that the phase compensation process
involved in the variable 'velocity of propagation region' results in the application of
an undesirable amplitude taper21 (emphasising the edges of the aperture). This
effect may be minimised by the choice of a suitable feed element to the lens to
counteract such a taper.
Circular arrays 303

Although there have been several experimental arrays, the practical utilisation
of beam cophasal circular arrays seems to have been mainly restricted to direction
finding applications and HF communication arrays where the availability of a
compact circular array antenna can replace a complete field of rhombic antennas.4
In such cases, wide bandwidth performance of the arrays is clearly essential and
scanning is performed by either a rotating coupler or set of switches for fixed
beams.

\ circular
array
elements

feed into
lens

circulator

Fig. 12.4 Use of a geodesic lens to feed a circular array

A different approach for electronic scanning for beam cophasal arrays employs
an array signal processing technique. This achieves a continuous fixed-rate beam
scanning by means of synthesising the correct form of phase variation (with time
and space) that needs to be applied to each element of a circular array. For the case
of beam cophasal patterns this corresponds to sinusoidal frequency modulation
and the feeds to each antenna element can be obtained by applying this frequency
modulated signal to a suitable tapped delay line. The above scheme has been
demonstrated for a receiving array by applying the same set of delayed FM signals
to act as local oscillators for each element of the array.11 The experimental system
employed 16 elements working at 400 MHz and a fixed scanning rate of 400 kHz.
A schematic diagram of this concept is shown in Fig. 12.5. The project was aimed
at the receiving system for a within-pulse scanning radar. The scheme can also be
used to rotate a restricted range of amplitude tapers on the array and the technique
can be extended for non-beam cophasal excitations provided that the desired phase
variation around the array is not discontinuous.
A form of continuous modulation scanning has also been adopted for an elec-
tronically scanned cylindrical array applied to a TACAN navigational aid.22*23 This
application requires a continuous rotation of a specific directional pattern.
Several of the array signal processing schemes discussed in Chapter 13 may also
304 Circular arrays

be applied to circular arrays. For example, if the output of a circular array is


sampled with a very fast switch (operating at the Nyquist rate for the bandwidth
of received signals) then the resulting samples may be fed to a matched filter which
can combine them together so that the single output represents a synthesised beam.
Multiple filters can thus generate multiple beams. This sort of concept is becoming
more attractive to implement in terms of digital technology.

sin(u>ot • ^ y 1 cos u> s t) i.f.amp.


equalising
#%)?osrcif"r
network to
produce 1Jy
frequency- u>s taper round
modulated array
local osc. to demodulator
and display

Fig. 12.5 Schematic of system for continuous electronic scanning of the directional pattern
of a circular array

Since the beam cophasal pattern of a circular array has only one main beam,
the beam pointing direction may be chosen to lie anywhere over the surface of a
hemisphere. This requires suitable phase (or time delay) excitation appropriate to
both the azimuth and elevation angle required. There is, however, no control of
the corresponding beamwidths in the two dimensions, which will be determined
solely by the angle of elevation of the beam. At first sight this seems attractive in
terms of generating one or more beams from a single ring array and then using
phasing or commutation to rotate the pattern. The main problem is that the vertical
beamwidth from a single ring is, in general, far too large and the need to apply
amplitude tapering makes it even larger.
A planar aperture, such as a circle, can, in principle, deflect a pencil beam over a
complete hemisphere (subject to mutual coupling and element pattern constraints).
A circular array can be regarded as the outer ring of such a planar aperture. (This
explains the high sidelobe performance of the beam cophasal pattern). The com-
putation of array phasing for 2-dimensional scanning of a circular planar array is
substantial though it is possible to adapt some of the array signal processing
schemes to this end.
Circular arrays 305

12.3 Amplitude-mode and phase-mode excitation of circular arrays

In the early 1960s workers in both the UK and USA started examining the direc-
tional properties of circular arrays in terms of the Fourier analysis of both the array
excitation and the corresponding far-field pattern24"26. This approach has proved
extremely attractive for synthesising directional patterns and has led to more
convenient methods of exciting arrays and of achieving electronic beam rotation.
The technique also offers the capability of compensating for the effects of mutual
coupling. It is, perhaps, surprising that the applications of this concept have been
slow to start but are now being applied in DF and null steering and this will be
discussed later. Before covering the details of such schemes, the basic theory of
phase modes and amplitude modes for circular arrays will be reviewed briefly.

12.3.1 Concept of phase and amplitude modes for continuous circular arrays

Phase modes: First consider the case of phase modes for continuous circular
arrays, where the effect of interelement spacing can be neglected (i.e. for very
small interelement spacing). The excitation function for any continuous circular
array is a periodic function of angle with period 2TT radians. The excitation function
may therefore be analysed in terms of a Fourier series. It is convenient to use the
complex form of Fourier series where the excitation of the array F(<t>) is expressed
as:
F{<t>) = Z C m exp(/m0) (12.5)
m=-N
Here Cm represents the complex Fourier coefficient of each spatial harmonic of the
array excitation and eqn. 12.5 represents the excitation function F(0) as the sum
of (27V-f 1) terms. Each term is called a phase mode excitation of the array. Any
arbitrary excitation function can be expressed in terms of these exponential phase
mode components. The mth phase mode consists of an excitation which is constant
in amplitude but with the phase varying linearly with angle so that there are m
complete cycles of phase change around 360° of angle. The corresponding phase
mode of order — m has the reverse direction of phase change with angle. The far-field
directional pattern arising from the mth phase mode turns out to be an identical
function having a similar phase variation with angle and a constant amplitude given
by a Bessel function coefficient. This mth far field phase mode may be represented
by:
2 7 r
f C ( / 0 ) [ ^ ( 0 ^ ) ] ^ 0 (12.6)
Z7
giving
A » W = C m / m / m (0r)exp(/m0) (12.7)
where j3 = 2TT/X.
Since each spatial Fourier harmonic of array excitation gives rise to a corres-
ponding spatial Fourier harmonic of the far-field directional pattern, the above
306 Circular arrays

argument for Fourier analysis can obviously also apply to the complete far-field
directional pattern which is also a repetitive function of angle. Therefore, it is
possible to take any desired far-field pattern and break it down into its complex
Fourier components and to separately excite each of these Fourier components
around the array, taking note of the appropriate Bessel function term which relates
the array excitation of that particular mode to the corresponding far-field signal
strength. Needless to say, the case of the zero-order mode is when all the elements
are excited in phase and this gives rise to an omnidirectional far-field pattern of
constant phase, as would be expected from symmetry. This corresponds to the DC
term in the time/frequency application of the Fourier series. The higher-order
modes are also omnidirectional but with higher rates of progressive phase change
with angle. This method of analysing and exciting such pattern from circular
arrays has many advantages for pattern syntheses.

Amplitude modes: If the cos/sin form of the Fourier series had been chosen, the
basic constituent pattern would have been m cycles of amplitude variation around
the array. This is termed an amplitude mode and gives rise to a similar amplitude
mode in the far-field pattern with the same Bessel function coefficient relating
these two functions. It can easily be seen that the summation of a pair of positive-
and negative-sequence phase modes of the same order will give rise to a corre-
sponding amplitude mode. By choosing the phase combination of each pair of
phase modes one can construct either a set of'cosine type' amplitude modes (used
for synthesising even-order patterns such as beam patterns) or 'sine type' amplitude
modes. The latter can be used to synthesise odd functions in directional patterns,
such as the monopulse-type split beam pattern. The far-field amplitude modes may
also be regarded as angular spatial harmonics.

Pattern synthesis: As an elementary example for synthesising a pattern of the


form sinNx/sinx, assume that (27V 4- 1) phase modes are excited on the array
according to the function:

This corresponds to exciting all the modes to the same level in the far field, resulting
in a directional pattern of the form:

W)= I C m / m / m (i3/-)exp(/m0) (12.9)


-N
giving
D(6) = I exp(/m0) = " ^ (12.10)
-JV sin t)
The beamwidth of this pattern is n/N.
This has resulted in the conventional sin Aft/sin x directional pattern for a
Circular arrays 307

uniform linear array, but it is important to realise that this function now extends
over 360° and relates to a 6 variable, rather than 180° and the sine 0 variable for
linear arrays.
The use of the Fourier technique for exciting circular arrays also means that
all the well known pattern synthesis techniques for linear arrays may be trans-
ferred to circular arrays. The taper function is then applied to the mode excitation
rather than to the aperture function. The resultant pattern is then identical to that
of the linear array theory, but corresponding to a 0 scale rather than a sin 6 scale.
In general, low sidelobe tapers will have higher excitations for the lower-order
modes. The types of network needed to excite such phase modes are discussed in
Section 12.3.4.

Choice of radius: The radius of an ideal continuous array has to be carefully


chosen to avoid conditions where one or more of the Bessel coefficients are zero.
If the radius is small, only a few modes can be excited, because the values of
Jm(fir) are all very small for m large. Physically, this means that the phase shift
round the array for the higher-order modes is so great that the radiated signals
effectively cancel in all directions. Excitation of modes much higher than fir corre-
sponds to superdirective conditions and, as with linear arrays, is only worth
attempting for small-sized arrays. The highest-order mode which can be excited at a
reasonable strength is therefore m ~ ± (fir). Using this as the value for TV in eqn.
12.10, the minimum beamwidth which can be obtained as

6 = n/fir = X/2r (12.13)

which equals the beamwidth of a linear array having the same dimensions as the
circular array diameter.
Even when the radius is not small, it is still possible for certain modes to be
difficult or impossible to excite if Jm(fir) is small, or zero. A physical explanation
of this effect can be obtained with the help of Fig. 12.6. For the direction 6 = 0,
the far-field phase of the signal received from a small arc of the continuous array
located at 0 and excited by mode m is

\p = ]3rcos0 + ra0 (12.14)


The maximum contribution to the far field comes from those parts of the circum-
ference where \jj is stationary, i.e. d^/dcp = 0. Thus there are two arcs for mode
m, with phase centres located at 0! and at IT — <pi where

sin0! = m/fir (12.15)


When m exceeds fir, sin 0X becomes greater than 1. In this case, the two phase
centres coalesce at 0 = TT/2, but there is little net radiation.
When r is large, a suitable discrete value of radius where /m(j3r)^=0 can be
obtained by noting that the asymptotic expansion of the Bessel function, valid
for the lower order modes (m < fir), is
308 Circular arrays

cos(pr-ir/4-mnl2)
J (12.16)
m\Pr) *
Clearly (3/* should be a multiple of n/2 to avoid zeros for any coefficient; i.e. the
diameter should be an integral number of half-wavelengths.

Fig. 12.6 Location of the 'phase centres' for the mth phase mode

The bandwidth of the array can be assessed, since |3r must not be allowed to
change by more than, say, 7r/8 to avoid excessive changes in the Bessel coefficients.
This condition gives
SX/A « ± 16X/r (12.17)

Vertical beamwidth: The radiation pattern in the vertical plane at B = 0 can be


obtained by substituting (cos a cos 0) for cos0 in eqn. 12.4, where a is the ele-
vation angle. The vertical pattern for the mth mode then becomes
Dm(ot) = C m / m / m ( ^ c o s a ) (12.18)
Thus each mode has a different vertical pattern. The zero order mode / 0 has its
peak at the zenith and as m increases, the peaks descend from the zenith. The
radiation in the horizontal plane is essentially 'sidelobes' of the Bessel functions.
The stability of the vertical pattern around a = 0 can again be assessed from the
Circular arrays 309

asymptotic expansions given above. If /3r cos a changes by 7r/8, there will be sub-
stantial changes in the Bessel functions. Thus the beam limits can be considered
to be given by
/3rcosa = /3r — TT/8 (12.19)
2
By approximating cos a as 1 — a /2, the vertical 'beamwidth' becomes
6 = 2a^^(XJ2r) (12.20)
which equals that of a linear endflre array of length equal to the diameter.
The above implies that the directional response with varying 6 but at angles out
of the plane of the array (a =£ 0) remains a phase mode but the amplitude of the
mode is a function of a. This represents a further problem for pattern synthesis,
since if the design produces a specified pattern in the plane of the circle, the shape
of the response at different vertical angles (different a) changes due to the different
relative levels of the phase mode components. A low sidelobe pattern at a = 0
could therefore deteriorate at higher angles. Fortunately, the above disadvantage
can be overcome by the choice of suitable directional elements as discussed later.

123.2 Phase modes in circular arrays employing a finite number of


omnidirectional elements
The concept of phase modes discussed in the previous Section relates to a con-
tinuous excitation function around the array. In practice, this is approximated
by means of a finite number of antenna elements. The use of discrete elements
does not change the basic concept. The array excitation is still a periodic function
of angle, so the concept of phase and amplitude modes still applies. In the case of
omnidirectional elements, these may be regarded as sampling the excitation func-
tion of a continuous array. By applying the sampling theorem it becomes apparent
that the circumferential sampling rate (i.e. the interelement spacing) should corre-
spond to at least twice the highest spatial frequency present in the array excitation.
This means that if the aperture excitation function contains harmonics up to a
maximum order of N, then the antenna array must contain at least 27V elements in
order to reproduce all these spatial harmonics.
As noted earlier, the maximum order mode that can be easily excited (without
superdirective excitation) is m = (fir). Therefore there should be at least (2j3r)
elements in the array, leading to a minimum element spacing of A/2. If a given
phase mode of order m is excited by means of a finite number of omnidirectional
elements the resultant far-field pattern may be computed by introducing a sampling
function S(0) which corresponds to a series of unit impulse functions at the ele-
ment locations. The function S(6) for omnidirectional elements may be expressed
in terms of its spatial spectrum as:

S(B) = £ ejnQd = 1 + 1 einQd + £ e~inqd (12.21)


Q=~°° Q-l Q=l

where n is the number of array elements.


310 Circular arrays

The aperture excitation for a phase mode of order m with discrete omnidirec-
tional elements is given by:
Fm(6) = CmeimeS(d) (12.22)
where Cm is the excitation level of the mode.
Fromeqns. 12.21 and 12.22:

Fm(6) = Cneime + C n £ (eiinq+m)d + e'Jinq'm)e) (12.23)

Since this represents Fm(6) in terms of an infinite series of phase mode excitations,
the corresponding far field can be expressed as:

+ I CmrgJg(Pr)e-J* + £ CmihJh(fr)eihB (12.24)


Q=l Q=l

where g = (nq — m), h = («g + ra), « — total number of elements and Cm = exci-
tation of rath mode.
The first term in the above equation represents the required far-field phase mode
pattern but the remaining terms correspond to higher-order distortion modes.
As in the case of the continuous array, there can be little radiation for modes
higher than ± 0r. Now the first higher-order distortion modes are separated by n
basic modes from the desired mode ra, so that it can be anticipated that the higher-
order modes due to the finite number of elements will not radiate very strongly
provided that
(n-m)>(3r (12.25)
Since the maximum value of m is |3/% this condition becomes
n>2$r (12.26)
which confirms that the inter-element spacing should be
d = 2nr\n<\\1 (12.27)
The properties of a circular array of omnidirectional elements with the above
spacing are thus similar to those of a continuous array. It may still be difficult to
excite all the desired modes and the array is narrowband. These difficulties can be
attributed to the two 'phase centres' of each mode, which are separated by a large
distance in a large-diameter array. This suggests that improvements can be obtained
by employing directional elements, which effectively suppress the radiation from
one of these phase centres.

12.3.3 Phase modes in circular arrays employing directional elements


The effect of directional elements in circular arrays has received only limited
attention, but a number of workers reported experimental results which indicated
Circular arrays 311

that they improved the performance of phase modes. These results showed that
the amplitude of the phase modes did not tend to zero in the horizontal plane
when/ m ($r) became zero.
The situation has been recently analysed by Rahim27'42 who has derived general
expressions for phase modes excited from directional elements and the following
represents extracts from his arguments.
Consider an ^-element array where each directional element has a pattern
G(\p) and the elements are disposed to retain circular symmetry. In most practical
applications this means directional elements pointing radially outwards from the
array. The pattern G(4J) can be represented in terms of its Fourier components:
P
G(i//) = £ Bhem (12.28)

where p is the maximum spatial harmonic of the element pattern.


Thus given G(i//), the coefficients Bk can be determined in the usual way by
Fourier analysis.
Assuming initially an infinite number of such directional elements (i.e. an
inter-element spacing of zero), then the expression for the far-field pattern of the
wth phase mode is:
1 r2Tr
Dm{B) = —Cm I G{6 — <t>) exp /[j3r cos (6—<f>) + m(j)]d(l) (12.29)

Combining eqns. 12.28 and 12.29,

Eqn. 12.30 can also be written as


!
ur\ //}\ Lr^ e JTTld \^ D I I es>J [(^M-fe)(0— d) + Bf COS (0— 9) ] 1a J / J L f\ \
m\v) ~ m L &k T~ JOL K9 ~~ V)
k=-p \27T I
(12.31)
Using the relationship
JL ff2Tr
j euy ( m 0 + ^ c o s < ^ (12.32)
2TT

the far-field equation becomes


w
Dm(6) = Cm\Am\enmu ^; (12.33)
where
(12.34)
Thus, for the case of zero interelement spacing, the function relating mode
312 Circular arrays

excitation on the array to the corresponding far-field pattern is no longer a single


Bessel coefficient but is now a sum of a finite series of such coefficients corre-
sponding to the Fourier coefficients of the element directional pattern. For most
practical elements this sum does not go to zero for any value of j3r, so that the
stability of mode amplitude due to changes of array radius or frequency is improved.
It is, however, interesting to note that directional elements of any form, but at
zero spacing, produce a pure form of phase mode without any distortion terms.
Section 12.3.2 discussed the effect of discrete elements and the effect of inter-
element spacing. This approach can be repeated for the above examples of direc-
tional elements, and this leads to a modified form of directional pattern given by:
Dm{6) = Cm\A
o

Cm £ {\Anq+m\expj[(nq +m)6

i// m - nQ ]} (12.35)
Higher-order distortion terms again arise due to the finite inter-element spacing.
Eqn. 12.35 also implies that distortion terms can be more significant with direc-
tional elements. This means that it is not possible to use directional elements in
order to relax the inter-element spacing, and hence economise on numbers of
elements. As noted earlier, there is little radiation from these higher-order terms
provided that the inter-element spacing is low (d < X/2) or the pattern does not
attempt to excite mode orders approaching the limit of (|3r).
Most experimental circular arrays with directional elements have employed slot
radiators or dipoles in front of reflecting screens. Examples of such arrays are
shown in Figs 12.19, 12.21 and 12.22, referred to later in the Chapter. Rahim
looked at the particular case of D(\p) = (1 4- cos \p) since this could be a realistic
approximation to many practical antenna elements. In this case the function
relating the far-field mode to the array excitation reduces to:
Dm{6) = Cmjme>m6{Jm($r)-jJ'm(Vr)) (12.36)
Examination of Bessel-function graphs reveals that when / m ($r) -> 0, J'm(fir) is
approximately a maximum and vice versa (except for $r -> 0). Therefore the resul-
tant far-field values of the modes becomes approximately constant. This offers
wide-bandwidth performance and the ability to design low sidelobe patterns by
means of accurate control of relative mode amplitude. It also explains why previous
experimental results have been better than expected.
An indication of the dramatic improvement achieved by the use of (1 4- cos \p)
elements can be seen from Fig. 12.7. This shows the theoretical variation of relative
mode gain with frequency for the first-order mode for both omnidirectional and
directional elements. These results correspond to an array having a diameter of 2X
at 300 MHz. Some experimental results28 confirming this for the same diameter
array are shown in Fig. 12.8. These results indicate that modes 0, 1 and 2 remain
stable in amplitude over an octave.
Circular arrays 313

The above evidence confirms the need for directional elements for circular
arrays where wide bandwidth and/or low sidelobes are required. Element patterns
of the form (1 + cos \j/) seem near optimum for such applications and the element
spacing should be A/2 or less to minimise distortion terms.

120 160 280 320 360 400


mode 1 frequency,MHz

Fig. 12.7 Theoretical results showing the stability in the amplitude of mode 1 vs. frequency
due to the use of directional elements of the form (7 + cos 0) for 2\ radius array at
300 MHz.
— '1 + cos 0' elements
omnidirectional elements

dB

mode 0
mode 1
°mode 2

200 225 250 275 300 325 350 375 400


frequency, MHz
Fig. 123 Experimental results for the variation of mode amplitude and frequency for a
2-wavelength circular array employing directional elements of the form {1 + cos 0)

The distortion terms associated with the element spacing give rise to departures
from true omnidirectional performance (amplitude ripples) and departure from the
linear variation of phase with angle.29 Fig. 12.9 shows a typical experimental result
for the polar diagram of a phase mode. These results relate to a 16-element array
314 Circular arrays

having element patterns of the from (1 + cos;//). The amplitude ripple is mainly
due to distortion terms corresponding to an element spacing of d = X. Fig. 12.10
shows a theoretical result for the variation of phase with angle for a first order
phase mode on a 4-element array. These cyclic phase errors are clearly important
for the design of DF systems and also represent phase errors for the constituent
components of any synthesised directional pattern.

Oth

Fig. 12.9 Experimental directional patterns showing the amplitude variation of the zero
and 1st order phase modes for a 2-wavelength circular array employing directional
elements of the form (/ + cos 0)

Rahim's analysis also showed that the use of directional elements of the form
(1 + cos \jj) also improved the vertical plane response of the phase modes27 by
removing the zeros of the pattern (which again corresponded to zeros of a Bessel
function.) Thus the use of a phase-mode synthesis technique in the plane of the
circle could retain the pattern shape out of this plane with such directional ele-
ments.

12.4 Excitation networks for phase modes on circular arrays

Several authors have shown that the well known Butler matrix network, used to
generate multiple beams for linear arrays, may also be used as a convenient way of
exciting the phase modes of a circular array. If an n x n port Butler matrix is
connected to an ft-element circular array, then each of its n input ports produces
the correct array excitation for the corresponding spatially-orthogonal set of n
possible phase modes that can be excited from such an array. This arrangement is
shown in Fig. 12.11. Excitation of each input port of the Butler matrix corresponds
to the excitation of a phase mode, so any (spatially band limited) pattern may be
excited from a circular array by applying the correct values of amplitude and phase
to these ports. This excitation function will depend upon the Fourier terms of the
180- far-field
phase shift
160 -

uo-
120-

100-

80-

60

40

20

0 —I . 1
180 210 240 270 300 330 360 9
-20
azimuth angle
-UO

-60-

-80

-100-

-120

-KO- §
oT
-160

-180
Fig. 12.10 Cyclic phase errors for first-order phase mode due to use of four omnidirectional elements at an interelement spacing of 0.4X
Calculated first order phase mode f o r four omnidirectional elements
Ideal first order phase mode (continuous array) Co
316 Circular arrays

desired directional pattern and the appropriate Bessel coefficient relating each
array phase mode to the far-field phase mode (which is also a function of the
directional elements used). This synthesises the desired directional pattern pointing
in one fixed direction.

T T T T T \
E E E E E E
i1 I2 i3 \U i5 in
Fig. 12.11 Butler matrix network with its output ports feeding the elements of a circular
array. Excitation of each input port to the network excites a corresponding
phase mode on the circular array
a = circular array with n elements
m = matrix network

A further important and useful property of this Butler matrix feed is that if a
linear progressive phase taper of value 0, <p, 2$, 3 0 , . . . is applied to the input ports
of the Butler matrix in addition to the above required excitation, then the resulting
signals to the array elements commutate around the circumference on the array.
This results in an angular rotation of the far-field pattern through the angle (j). This
makes the Butler matrix feed network a very attractive system for both excitation
and electronic rotation of patterns for circular arrays. A schematic representation
of this system is shown in Fig. 12.12. This result was shown by Davies26 and
experimental arrays based upon this principle have been described by Sheleg10
and by Chow and Davies30
Fig. 12.13 shows an experimental result from Reference 30 which relates to the
above type of scanning system when feeding a 16-element circular array of mono-
pole elements located over a flat ground plane at 400 MHz. The excitation involved
amplitude modes of order 0 to 6 with an amplitude taper to achieve — 23 dB side-
lobes. The results relate to a continuous beam scanning at a rate of 100 kHz.
An alternative way of looking at this Butler matrix feed system is to note that
it performs a transformation between the excitation techniques of linear and
circular arrays. Clearly, any technique used to achieve linear progressive phase shifts
across a linear array may be used with a Butler matrix to achieve beam deflection
Circular arrays 317
for circular arrays. The results of Fig. 12.13 were obtained from a single-sideband
type array signal processing system designed for continuous scanning for linear
arrays. This scanning concept is discussed in Chapter 13.

Fig. 12.12 Schematic diagram of arrangement for electronic angular deflection of any
directional pattern formed by a circular antenna array. The pattern is excited
in terms of its constituent phase modes via the Butler matrix and a linear pro-
gression of phase shifts across the input ports of the matrix produces linear
angular shifts of the resultant directional pattern
a = phase shifts to deflect pattern through angle 0
b — values of amplitude and phase shift required to synthesise directional pattern
m = matrix network

Fig. 12.13 Experimental directional pattern of 16-element circular array undergoing con-
tinuous electronic pattern rotation at 100kHz (— 23 dB sidelobe level)

In principle, all designs of multiple-beam networks used for linear arrays may
also be employed to excite phase modes of circular arrays. The phase mode exci-
tation, however, requires constant values of phase shift over the frequency band of
interest. This can only be achieved by the Butler matrix configuration (when
suitably modified for wideband operation) since other forms of network, such as
Maxson matrix and lens configurations, are based upon path length (time delay)
compensation. Wide-bandwidth Butler matrix networks have been constructed for
16-port designs28 and the technique is, in principle, extendable to larger systems31
at the expense of additional complication. A recent design9 for an octave bandwidth
318 Circular arrays

DF system is described in Section 12.6.3. It makes use of a wideband Butler con-


figuration and a 16-element circular array.
The cost and complication of designing Butler matrix networks increases rapidly
with the number of elements. Designs become progressively difficult for more than
32 elements. This means that the above feed system tends to become complex for
apertures greater than about 5X diameter. For receive-only configurations, IF pro-
cessing, digital processing and array signal processing can be used to overcome this
limitation. Despite this, the scale of complication increases significantly for larger
radius arrays.

12.5 Mutual coupling in circular arrays

All arrays suffer from the effects of mutual coupling between their constituent
elements and circular arrays are no exception. In general, the effects are more
severe for such arrays since significant coupling can occur between elements located
diametrically opposite each other in addition to the strong coupling to adjacent
element locations. However, in the case of the excitation of a single-phase mode
around the array, the electrical environment of all the elements is identical, due to
symmetry. This means that all the elements will exhibit an identical radiation
impedance. Nevertheless, mutual coupling effects still modify the value of this
radiation impedance, but in the case of a single-phase mode, all the values are
changed by the same amount.32 Such a radiation impedance, under phase-mode
excitation, is often referred to as the phase-sequence impedance of that mode.
An identical argument applies to all the other phase modes so that the effect
of mutual coupling is to modify the various sequence impedances and hence the
matching at the input port of the matrix associated with each phase mode. There-
fore if phase mode excitation is employed with a network (such as a Butler matrix)
having a separate feed port for each phase mode, it is possible to completely com-
pensate for the effects of mutual coupling by compensating for this impedance
mismatch at each input port. Thus a set of fixed impedance compensations at
each input port will be adequate to compensate for the effects of all mutual
coupling, irrespective of the excitation of the array or of the direction of beam
pointing. For wide-bandwidth operation it is necessary to match each mode impe-
dance over the operating band. The use of directional elements serves to reduce
the effects of mutual coupling since it can reduce diametrical coupling across
the array. The above compensation can, however, still be applied.
Correction of mutual coupling effects may be performed theoretically by
calculating the value of the phase-sequence impedance, and hence calculating the
change required in the excitation of the mode. Reference 32 gives detailed calcu-
lations for such mutual coupling. Alternatively, it is possible to obtain such results
experimentally by exciting each phase mode one at a time, measuring the field
strength and phase achieved in the far field and then modifying the excitation
of the appropriate port in order to achieve the desired value of field strength for
Circular arrays 319

each constituent mode. This attractive feature gives powerful advantage to circular
arrays for those applications where mutual coupling can limit the performance of
an antenna. This can apply, in particular, to low sidelobe applications and small
aperture arrays.
In the case of wideband excitation, in principle, it is necessary to produce an
excitation for each mode which varies with frequency in such a way as to maintain
the correct pattern, taking account of mutual coupling. There is little published
work on this topic but the limited evidence on the wideband stability of modes
when using directional elements suggests that this may not be impracticable.

12.6 Applications of circular arrays

The rather limited number of applications for circular arrays has been mainly due
to the problems and complications of array excitation and beam rotation. In the
case of beam cophasal patterns this has been particularly difficult and, in general,
has been solved by multiple beams or mechanical rotating goniometers. The
development of phase mode theory in the 1960s represented a substantial step
forward, coupled with the use of Butler matrix systems to feed the modes. Antenna
engineers have studied these concepts for several applications including radar, null
steering and DF systems.

12.6.1 Communication applications


Circular arrays have been used for many years in the HF band for both communi-
cation and direction finding.3 The systems employ beam cophasal excitation with
time delay compensation to achieve broad bandwidth operation. The elements are
usually a monopole or a combination of driven and parasitic elements. Multi-octave
performance can also lead to two separate circular arrays covering different parts
of the band. Such Wullenweber arrays usually excite an arc of about 120° and the
active arc is rotated by means of a goniometer involving capacitive coupling
between plates on a stator and a rotor. Networks to achieve the phase delays and
any amplitude taper are then located on the rotor and the signals brought out via
slip rings.
A different approach for wide-bandwidth arrays is to form multiple fixed beams
from the elements. Longstaff and Davies19 discuss various configurations for such
beams and suggest ways to vary the active arc of excitation over the frequency band
of interest in order to retain a fan beam coverage in the vertical plane. Experimental
results were provided on a model array of monopole elements covering 200-
1400 MHz. Fig. 12.14 shows some experimental results in this system.
Starbuck4 described an operational HF circular array designed and manufactured
by the Plessey Company and covering the band 1.5-30 MHz. This antenna is based
upon the concept of 24 fixed beams formed by time delay circuits, covering 360°
of azimuth. It employs two concentric circular arrays of elevated feed monopoles
to increase the overall bandwidth. A wide-bandwidth circular antenna based upon
320 Circular arrays

multiple input feeds is the HF wire grid Luneberg lens.21 An antenna based on this
principle was constructed for the FAA near Honolulu33 covering the 3-30 MHz
band and consisting of a 600 ft-diameter wire grid structure.

-180--150"-120''-9Cf-60'-3Cf 0 30" 60" 90 120 150 180


a z i m u t h angle •

Fig. 12.14 Some measured directional patterns of the 16-element wide-bandwidth circular
array described in Reference 19. Curve (a): 205MHz; curve (b): 480MHz;curve
(c): 1350 MHz

An application of phase mode excited circular arrays relates to null steering


arrays for communications. Here a single-phase mode is used to provide omni-
directional coverage and a second (adjacent order) phase mode is introduced to
provide a single directional null. If the two phase modes have equal amplitudes in
the far field there will always be one angle at which the signals from the two modes
are in phase opposition. The combined signal will therefore exhibit a cardioid
(single null) pattern. By changing the phase shift to either mode the angle of the
null can be steered through 360° by means of a single phase shifter.8
This configuration is shown in Fig. 12.15 for a simple 4-element array. Fig.
12.16 shows a photograph of an experimental 4-element array operating at 150 MHz
and employing dipole elements located around a central support mast. Fig. 12.17
shows an example of a typical null pattern. Additional phase modes may be used to
either reduce the null width of the pattern or to introduce additional nulls.8*34'35
Computer control can also be included for automatic null steering.29 Since an
^-element circular array is capable of generating n independent phase modes, this
leads to a capability of generating (n — 1) independent directional zeros in its
directional pattern.
The above concept can be extended to communication arrays capable of oper-
ating over broad bandwidths and with the facility of introducing directional beams
to enhance gain in a required direction plus the capability of employing nulls to
counteract unwanted directional interference or jamming.
The ability to compensate for mutual coupling in circular arrays is important in
several of the above applications and also opens the way to interesting developments
for very small radius circular arrays. In such cases, the small radius arrays (less than
Circular arrays 321
0.1 A radius) may be used to form directional beams or null steering patterns.
Experimental results for a small radius null steering circular array have been des-
cribed36 for a radius of 0.04X. These are really super-directive type antennas, but
one of the real problems of super-directivity, namely mutual coupling, can be
overcome in this instance due to the symmetry. There is, nevertheless, still a
matching problem for such antennas which usually restricts their operation to
narrow bandwidths.

> [first-order
zeroth mode
^ Y phase mode
variable [~7| null-setting
phase-shifter A ^ o . L_jJ attenuator

(c)

Fig. 12.15 Schematic diagram for broadband feed system for producing a single steerable
directional null

The siting problems of omnidirectional antennas on broadcast masts or ship


superstructures can lead to shadowing due to the supporting structure. Circular
or cylindrical wrap-around arrays have therefore sometimes been adopted.5 Wyatt
has described a simple 4-element circular arrays employing four broadband dipoles
used as a UHF wrap-around antenna for naval communications.
The same concept can also be applied to satellite antennas or military platforms
such as remotely-piloted vehicles. The use of a circular array merely to achieve
omnidirectional coverage seems rather complex in such cases. This had led to the
examination of combining this feature with a degree of null steering29 or beam
forming from the same array. This concept could be attractive for shipborne
antennas.

12.6.2 Radar applications


It is, perhaps, not surprising that circular arrays have not yet been applied in
operational radar systems since the extent of application of electronically-steered
322 Circular arrays

linear arrays is still quite limited. Nevertheless, several experimental systems and
subsystems have been constructed. These have included an array employing diode
switching of signals, constructed by Wheeler laboratories, and a further switching
scheme which could also rotate an amplitude taper, developed by the Naval Elec-
tronics Laboratory at San Diego. Such switching systems have tended to be com-
plex, particularly where amplitude tapers have been involved.

Fig. 12.16 Photograph of experimental 4-element null steering array (see Reference 8)

Sheleg,10 at the Naval Research Laboratory, Washington, has described an


experimental 32-element circular array (5.1 X diameter) fed from a Butler matrix
and operating at 900 MHz. The system included phase shifters at the mode inputs
to control beam direction. A photograph of this array is shown in Fig. 12.18. A
subsequent array built by NRL involved a cylindrical structure consisting of 32
vertical linear arrays (each of 8 elements) disposed around a cylinder (see Fig.
12.19). This system employed phase modes for the circular array (horizontal
plane), excitation and frequency scanning in elevation. Both sum and difference
beam monopulse patterns could be formed and scanned in two dimensions. A set
of beam patterns covering 360° azimuth for 40° elevation is shown in Fig. 12.20.
Two different circular arrays with electronic scanning were constructed at the
University of Birmingham, UK. The first14 employed fast (400 MHz) scanning of a
Circular arrays 323
beam cophasal receive-only pattern from a 16-element array of monopoles. The
schematic system is shown in Fig. 12.5 and was mentioned in Section 12.2.2. This
concept can only produce a continuous-rate beam rotation (and not beam posi-
tioning). Although the system is simpler than many competing schemes, it is
difficult to rotate an amplitude taper without more complication. The second
scheme30 employed omnidirectional elements plus a Butler matrix feed. The
system employed single-sideband modulation of the input to the phase modes to
achieve continuous rotation of the pattern as mentioned in Section 12.4. One of
the directional patterns is shown in Fig. 12.13 undergoing continuous rotation at
100 kHz.

(/ \ X /^n-A /
-10-
/ \ —A./ >4—null width
r58
dB \ X / °
\ f-i
-20-
circular l ^ \ I \ 1
array \ / \ /

-30-

I
0° 90° 180° 270°
angular positions of array elements

Fig. 12.17 Experimental null patterns for 4-element null steering array at 153 MHz. Curve I
shows null pointing in directions corresponding to element locations. Curve II
shows null pointing between direction of element locations

Experimental circular array systems have also been constructed using lens
techniques to feed the array. These, in general, provide beam cophasal type exci-
tation. Lens systems may employ geodesic lens configurations or an R-kR lens. This
latter approach has been adopted by the Naval Electronics Laboratory, San Diego,
employing a parallel plate construction for the lens. Papers covering several such
systems were presented in two conferences on conformal arrays37'38 at NEL San
Diego in 1970 and 1972.

12.63 Circular arrays in direction finding


The simplest example of a circular array used for DF applications is probably the
4- or 5-element array used for cathode-ray direction finding systems (CRDF). Four
dipole elements form a square and are excited as two pairs (i.e. two Adcock cou-
plets). A single dipole (or monopole) located in the centre of the square is used as a
phase reference to remove the 180° ambiguity of angle of arrival. These systems
are used to provide instantaneous display to aircraft controllers of the bearing of
an aircraft via the VHF voice link.
A significant development of DF systems was the 'commutated aerial direction
324 Circular arrays

fmder' whose principles were first described by Earp and Godfrey.39 This involved
a circular array of antenna elements whose outputs were sequentially sampled by a
switch. Early versions used a mechanically rotating switch, but electronic switching
was subsequently used.

Fig. 12.18 Photograph of experimental 32-element circular array at NRL, Washington

The output from such a switch, when the array is excited by a single plane
wave, represents a sampled version of a phase-modulated waveform in which the
phase deviation is given by the diameter of the array in wavelengths. By comparing
the phase of the phase modulation with that of the switching control, the angle
of arrival may be determined. Increased bearing accuracy can be obtained from
larger diameter arrays without the need for rotating parts. This technique used
in reverse for transmission is also the basis for the Doppler VOR and Tacan
antennas.22'23
Wullenweber arrays have been extensively used for DF applications in the HF
band and this work is reviewed by Gething.3'40 Several large antennas of this type
have been constructed in various parts of the world. The University of Illinois
system41 covers the 4-16 MHz band employing a single ring of 120 elements on a
circular of diameter 955 ft.
More recently work has been reported on octave-wide microwave DF by the
use of a circular array comprising 16 elements fed by a wide-bandwidth Butler
matrix.9 This provides a DF accuracy of 0.17° rms over the 2-4 GHz band. The
principle involves comparing the phase of signals received by two phase mode
1
to
O
Fig. 12.19 Photograph of NRL cylindrical array employing printed-circuit dipoles and feeds for the linear arrays forming the cylinder
to

Fig. 12.20 Experimental directional patterns for the cylindrical array of Fig. 12.19. The set of patterns correspond to 36Cf coverage in azimuth
when the elevation beam direction is 4<f above horizontal
Circular arrays 327

patterns. If the order of the modes differs by one, then the phase difference gives
an unambiguous output of bearing. Higher accuracy can be achieved by using
more than two modes. This concept is being marketed by Anaren for an ESM
system. A photograph of the antenna is shown in Fig. 12.21.

Fig. 12.21 Photograph of 16-element circular array used for ESM direction finding system
by Anaren

The use of phase modes in this type of DF system emphasises the wide-bandwidth
properties of the phase-mode concept when used with suitable directional elements.
The accuracy of such a system is also affected by the higher order distortion terms.
Fig. 12.10 showed a theoretical example of phase errors for a phase mode pattern
due to the finite interelement spacing. This will translate into a bearing error for
this type of DF system. The solution is to employ many elements and use low-order
modes or to reduce the interelement spacing, or simply to 'calibrate' the mode's
phase versus angle response.

12.7 References

1 CROMPTON, C: 'Navalradio direction-finding',/. IEE, 94, Pt. IIIA, 1947, p. 132


2. CARTER, P. S.: 'Antenna arrays around cylinders', Proc. IRE, 3, Dec. 1943, p. 671
328 Circular arrays

3 GETHING,P. J. D.: 'High-frequency direction finding', Proc. IEE, 113, Jan. 1966, pp.
49-61
4 STARBUCK, J. T.: 'A multiple-beam HF receiving aerial system', Radio Electron Engr.,
37, April 1969, pp. 229-235
5 WYATT, J. E. G.: The design of an omnidirectional UHF wrap-round antenna', /. Royal
Naval Scientific Service, 29, pp. 251-258
6 HANSEL, P. G.: 'Doppler-effect omnirange', Proc. IRE, 41, Dec. 1953, p. 1750
7 GREGORWICH, W. S.: 'An electronically despun array flush-mounted on a cylindrical
spacecraft', IEEE Trans., AP-22, Jan. 1974, pp. 71-74
8 DAVIES, D. E. N., and RIZK, M. S. A. S.: 'A broadband experimental null-steering
antenna system for mobile communications', Radio Electron Engr., 48, Oct. 1978, pp.
511-517
9 REHNMARK, S.: 'Improved angular discrimination for digital ESM systems', Proc. MM80,.
Microwave Exhibitions and Publishers Ltd., 1980
10 SHELEG, B.: 'A matrix-fed circular array for continuous scanning', Proc. IEEE, 56, Nov.
1968,pp. 2016-2027
11 FENBY, R. G., and DAVIES, D. E. N.: 'Circular array providing fast 360° electronic beam
rotation', Proc. IEE, 115, Jan. 1968, pp. 78-86
12 SHELEG, B., and WRIGHT, B. D.: 'A 3-D electronically scanned S-band cylindrical
array', Proc. of conformal array antenna conf., NELC, San Diego, CA. Jan. 1970
13 BOGNER, B. F.: 'Circularly symmetric RF commutator for cylindrical phased arrays',
IEEE Trans., AP-22, Jan. 1974, pp. 78-81
14 SKAHILL, G., and WHITE, W. D.: 'A new technique for feeding a cylindrical array',IEEE
Trans., AP-23, March 1975, p. 253
15 ZIEHM, G.: 'Optimum directional pattern synthesis of circular arrays', Radio Electron
Engr., 28, Nov. 1964, pp. 341-355
16 FENBY, R. G.: 'Limitations on directional patterns of phase compensated circular arrays',
Radio Electron Eng., 30, Oct. 1965, pp. 206-222
17 JAMES, P. W.: 'Polar patterns of phase corrected circular arrays', Proc. IEE, 112, 1965,
p.1839
18 WATANABE, F., GOTO, N., NAGAYAMA, A., and YASHIDA, G.: 'A pattern synthesis
of circular arrays by phase adjustment', IEEE Trans., AP-28, Nov. 1980, p. 857
19 LONGSTAFF, I. D., and DAVIES, D. E. N.: 'A wideband circular array for HF communi-
cations', Radio Electron Engr., 35, June 1968, pp. 321-327
20 HOLLEY, A. E., DUFORT, E. C, and DELL-IMAGINE, R. A.: 'An electronically scanned
beacon antenna', IEEE Trans., AP-22, Jan. 1974, p. 3
21 TANNER, R. L., and ANDREASON, M. G.: 'A wire-grid lens antenna of wide application',
Proc. IRE Trans. PGAP, July 1962, p. 408
22 CHRISTOPHER, E. J.: 'Electronically scanned TACAN antenna', IEEE Trans., AP-22,
Jan. 1974, p. 12
23 SHESTAG, L. N.: 'A cylindrical array for the TACAN system', IEEE Trans., AP-22, Jan.
1974, p. 17
24 TILLMAN, J. D., HICKMAN, C. E., and NEFF, H. P.: The theory of a single ring circular
array', Trans. Amer. Inst. Elec. Engrs., 80, Pt. 1,1961, p. 110
25 LONGSTAFF, I. D., CHOW, P. E. K., and DAVIES, D. E. N.: 'Directional properties of
circular arrays', Proc. IEE, 114, June 1967, pp. 717-718
26 DAVIES, D. E. N.: 'A transformation between the phasing techniques required for linear
and circular aerial arrays', Proc. IEE, 112, Nov. 1965, pp. 2041-2045
27 RAHIM, T.: 'Directional pattern synthesis in circular arrays of directional antennas', Ph.D.
Thesis, Univ. of London, 1980
28 RAHIM, T., GUY, J. R. F., and DAVIES, D. E. N.: 'A wideband UHF circular array',
Proc. 2nd International Conference on Antennas and Propagation, April 1981
29 DAVIES, D. E. N., and GUY, J. R. F.: 'A UHF communication antenna employing
Circular arrays 329

open-loop control null steering*, Joint International IEEE/AP-S Symposium, National


Radio Science Meeting and International IEEE/MTT Symposium on Antennas and Pro-
pagation, 15-19 June 1981
30 DAVIES,D. E. N., and CHOW, P. E. K.: 'An experimental array with electronic beam
rotation', Proc. European Microwave Conference, 1969
31 WITHERS, M. J.: 'Frequency-insensitive phase-shift networks and their use in a wide-
bandwidth Butler matrix', Electron. Lett., 5, Oct. 1969, pp. 496-497
32 KING,R. W. P., MACK, R. B., and SANDLER, S. S.: 'Arrays of cylindrical dipoles'.
Cambridge Univ. Press, 1968
33 ROGERS, G. V.: 'Lens-like antenna: low noise, less space', Electronics, 28 Feb. 1964,
p. 25
34 DAVIES, D. E. N. and RIZK, M. S. A. S.: 'Electronic steering of multiple-nulls for circular
arrays', Electron. Lett., 13, Oct. 1977, pp. 669-670
35 LIM, J. C: 'Introduction of a sharp steerable null response in an otherwise omnidirectional
pattern using a circular array', Radio Electron Engr., 47, Jan./Feb. 1977, pp. 30-32
36 DAVIES, D. E. N., and RIZK, M. A. A. S.: 'A small radius circular array with 360° null-
steering capability'. IEE Conf. Publ. 169, Antennas and Propagation, 1978, p. 60-64
37 Proc. Conf. on Conformal Array Antennas. Naval Electronics Lab., San Diego, Calif.
Report No. TD95,13-15 Jan. 1970
38 Proc. Array Antenna Conf. Naval Electronics Lab., San Diego, Calif. Report No. 155,
Vol. 1, Pts. 1 and 2, 22-24 Feb. 1972
39 EARP, C. W., and GODFREY, R. M.: 'Radio direction finding by the cyclical differential
measurement of phase', Proc. IEE, 94, Pt. IIIA, 1947
40 GETHING, P. J. D.: 'Radio direction finding', Peter Peregrinus, IEE Electromagnetic
Series 4,1978
41 HAYDEN, E. C: 'Propagation studies using direction finding techniques', /. Res. Nat. Bur.
Stand.,65D, 1961, p. 197
42 RAHIM, T. and DAVIES, D. E. N.: 'Effect of directional elements on the directional
response of circular arrays', Proc. IEE Vol. 129 Pt H No. 1 Feb. 1982. p. 18-22.
Chapter 13

Array signal processing


D. E. N. Davies, K. G. Corless, D. S. Hicks, K. Milne

13.1 Introduction

Array signal processing (ASP) represents a class of antenna system where the output
from an array is subjected to various forms of signal processing (e.g. phase and
amplitude control, time delay or frequency translation) in order to produce outputs
which can simultaneously provide angular information relating to several different
directions.
When the outputs of a multiple element array are combined via some passive
phasing network such as a corporate feed or combination of hybrids, the phasing
will usually arrange for the output of all the elements to add coherently for a given
(main beam) direction. It is interesting to consider what happens to the power asso-
ciated with an incident wave arriving from an angle corresponding to a zero of the
directional pattern. If the phasing network is constructed from passive transmission
lines, then such incident power must be re-radiated. If the phasing network contains
hybrids, then this incident power may flow into the 4th port terminations of such
hybrid units. This is an indication that, although the array receives signals from a
wide range of incident angles, the phasing network provides the directional res-
ponse. Clearly, it is possible to utilise this power by designing multiple port phasing
networks which can simultaneously provide several outputs corresponding to differ-
ent beam directions. This aspect has already been discussed in Chapter 6 which also
drew attention to the need for the directional patterns resulting from such multiple
beams to be spatially orthogonal in order that the outputs from these ports may be
electrically uncoupled.
This corresponds to a simple example of an ASP antenna system which could be
used to transmit or receive independent signals to or from different directions. In
this instance, the processing consists of time delays or phase shifts needed to
synthesise multiple beams, but several other techniques including modulation
methods may be employed to control the shape and pointing direction of the
antenna directional characteristic. In some instances the processing schemes may
make use of non-linear techniques (e.g. signal multiplication, square-law detection,
hard limiting etc.).
Array signal processing 331

The principal use of ASP techniques has so far been in radar, sonar, radio astron-
omy and certain navigational aids. In such cases, the properties of the directional
pattern are determined by the signal processing system which is also closely related
to the system application. It is therefore necessary to discuss ASP techniques in
relation to some form of radar or other application area, since the 'antenna system'
can no longer be considered as a separate isolated unit.
A different aspect of ASP which has received considerable attention over the last
few years is adaptive arrays and null steering antennas. Here, instead of defining a
number of wanted directions for receiving signals we can specify some wanted and
some unwanted directions; the latter may contain high-level interfering signals
which must be minimised by means of directional nulls in the patterns. In some
cases the directions of these nulls may be fixed or pre-programmed, but in many
cases the antenna response must be adaptive. This means that the shape of the
directional pattern is determined in response to the angular distribution of received
signals in order to try to meet some desirable criterion.
The simplest form of adaptive and null steering antenna is probably the sidelobe
canceller. This can steer a single directional null within the sidelobe region of a nor-
mal directional pattern. Multiple null adaptive arrays in general define a 'wanted
direction' and then introduce one or more feedback controls to minimise the signals
from all other directions. Null steering systems have applications in radar, sonar and
communication systems. In the latter case, there may be additional complications if
it is not possible to specify the direction of the wanted signals. The field of null
steering and adaptive arrays is covered in Sections 13.6 and 13.7.

13.1.1 Applications of array signal Processing


One of the earliest examples of an ASP system was in the communications field in
the form of a multiple element antenna system for HF communications,1 Here, the
output from a number of randomly located HF antennas could be phased and com-
bined together to form one or more beams pointing in different directions. The
beam forming was performed after amplification to remove coupling between the
beams. More recent communications examples include multiple beam antennas for
TV broadcasting by satellite which can provide a number of switched, contoured
beams to provide coverage of different geographical locations.
Radio astronomy and radar astronomy have made considerable use of ASP tech-
niques. Perhaps the simplest example is the use of mterferometric arrays to deter-
mine angle measurements of radio sources. Radio astronomy also saw the invention
of the synthetic aperture concept by Ryle.2 Here, aperture synthesis is used in a
post-demodulation sense. The outputs of two antenna elements are multiplied
together and the resulting signal recorded. The locations of the elements are then
changed in such a way as to reconstruct the component cross products of the sig-
nals which would have been received from a large fixed aperture antenna. Aperture
synthesis has also been applied in radar astronomy by employing a coherent radar
source and taking account of the movement of the antenna due to the earth's move-
ment, to synthesise an aperture corresponding to the movement of the antenna.
332 Array signal processing
This aperture synthesis technique is also used to achieve very high resolution coher-
ent radars on mobile platforms (aircraft and spacecraft) by storing and processing
the signals received over a period of time while the platform moves a distance cor-
responding to the length of the synthesised aperture.
Many surveillance sonar systems employ ASP techniques of necessity due to the
slow velocity of propagation in the medium. The delay between transmitted and
received signals is often some tens of seconds so that the data renewal rate for a
single scanning beam would therefore be several minutes. This is likely to be unac-
ceptable for most operational systems. Many sonars therefore employ floodlight
transmitters with either multiple receiving beams or Nyquist rate scanning of the
receiving beam.3 Sonar also provides an interesting application for the use of fast
Fourier transform programs to perform angular spectrum analysis.
There are several examples of ASP techniques in the navigational aid field such as
the Doppler VOR beacon and recent work on microwave Doppler landing systems.4
In both examples a transmitter is connected to a number of antenna elements in
time sequence so that the phase centre of the radiation moves along the linear array
(for the Doppler microwave landing system) or around a circular array (for the Doppler
VOR). The resulting radiation produces a spatially coded signal used to indicate
direction.
The most extensive development of ASP techniques can be found in the radar
field. The simplest example is the monopulse radar incorporating 2 beams (or 4
beams for 2 dimensional monopulse) to provide angular information by either
amplitude comparison or phase comparison techniques on a single received pulse.
This technique was subsequently developed to form stacked beam 3D radars. Such
radars generally incorporate a vertical stack of feed horns and either a reflector or
lens system, thus forming a number of beams in elevation. The whole antenna
rotates and each beam feeds a separate transmitter/receiver channel. Some stacked
beam systems employed a single fan transmitter (i.e. floodlight illumination in ele-
vation) with a stack of receiving beams. Later radars employed different transmitter
frequencies for each beam. Recent alternatives have involved the use of Nyquist
rate electronic scanning in the vertical plane5 or multiple beam networks.6
A particularly interesting radar involving advanced signal processing in both
range and angle is the AR 3D radar incorporating the so-called 'Squirt' tech-
nique.7 ' 8 This employs a frequency modulated transmitter with a dispersive end-fed
linear antenna array having an artificially enhanced frequency versus angle charac-
teristic and a within-pulse compression equalisation over the whole of the trans-
mitted bandwidth. In this system non-linear frequency modulation is used to pro-
vide a tailored coverage diagram.
The range of application of ASP techniques is increasing, primarily due to
improvements in the power and flexibility of modern signal processing technology.

13.2 ASP receiving systems for radar

13.2.1 Multiple beams


Although ASP techniques have been used in various applications, we shall take an
Array signal processing 333
example of a simple radar system providing a one dimensional sector coverage from
a linear array in order to illustrate a range of ASP techniques, before going on to
more general applications. Fig. 13.1 shows a sector which is illuminated by a broad
beam floodlight transmission and covered, on reception, by a group of n narrow
fixed receiving beams. If the operation of this type of radar is compared with the
more conventional arrangement of a single narrow common T & R beam pointing
to each of n separate directions in time sequence, it is evident that, despite the
reduced gain of the transmitting antenna, the average power illuminating each
target is the same in both cases (when averaged over a complete cycle of the scan-
ning radar). If the peak power per pulse received by a given target is P for the scan-
ning radar, the corresponding figure for the floodlight transmission will be Pjn. The
target, however, will now be illuminated n times as frequently by the floodlight
transmitter.

sector illuminated
by wide-beam floodlight
transmission

n separate receiving
beams covering sector

Fig. 13.1 Array signal processing radar coverage.


A sector is illuminated by a wide-beam floodlight transmission and n separate over-
lapping beams are employed on reception.

The use of multiple receiving beams with a single floodlight transmission in this
configuration has a number of important consequences. First of all, there is a sub-
stantial increase of data renewal rate, since information is now received from all tar-
gets at all ranges and all bearings for each transmitted pulse. The essence of this is
that the ASP array effectively looks in many directions simultaneously. It is this
feature which essentially distinguishes ASP from a conventional electronic beam
scanning system. An amplitude-comparison monopulse antenna may therefore be
regarded as a 2-beam form of ASP.
The effect of the above processing on the maximum range performance of the
radar is rather complex, since in practice it will depend upon the form of pulse to
pulse integration used. This is because the ASP radar receives N times as many
pulses, each having one Nth of the power of the conventional, sequential scanning
radar. Other features of such a radar include the use of fixed antennas so that the
334 Array signal processing
system may be classed as an electronic scanning system, but without the need for
electronic phase control of the high power radar transmissions. The fixed antennas
also give a certain advantage for coherent radars and MTI systems by the complete
elimination of loss and spectrum spreading effects due to scanning modulation (this
is an advantage which does not apply to conventional electronic scanning). An
important feature of such a radar is that the overall sidelobe performance of the
system is determined entirely by the one-way sidelobes of the receiving array. Since
sidelobe levels are inevitably determined by the phase and amplitude accuracy of
the aperture excitation function, such accuracy must also be reflected in the cor-
responding receiver signal processing.
The simplest way of implementing such a multiple beam receiver would be to
employ a reflector or lens incorporating a number of primary feeds. Alternatively,
it is possible to employ a multiple beam matrix network such as the Butler matrix
or the Blass matrix. It would be necessary to provide a separate receiver at each out-
put port as shown in Fig. 13.2, but the directional patterns of such outputs are
necessarily limited by the spatial orthogonality criterion discussed earlier. The
antenna gain of each beam will be equal to the full gain of the array in each direc-
tion (taking into account the cos# variation of effective aperture). In practice, this
gain would be reduced by any loss associated with the multiple beam matrix or
lens/reflector system.

• •
T
Rx Tx

beam-forming network
0
1 ! I
Fig. 13.2 Schematic arrangement associated with the coverage described in Fig. 1
Antenna array feeding multiple beam forming network on reception

If band limiting RF amplifiers are introduced after each antenna element, this
allows the SNR of the system to be defined before any beam forming or processing
is introduced. This offers additional flexibility and opens the way to many new pro-
cessing schemes. The first additional flexibility is that the multiple beam network
may be lossy and may produce non-orthogonal beams without restriction on either
antenna gain, sidelobe level or beam overlap. It may also be more convenient to fre-
quency change the output of all the elements down to some intermediate frequency
where the design of the multiple beam matrix network may involve more conven-
ient electronic components.
Array signal processing 335

It is also worth noting that amplitude tapers applied after amplification will
obviously taper both signals and receiver noise together and hence not worsen SNR
to such a large extent. In practice, this is worth about 1-2 dB in SNR.
Multiple beam networks may also be regarded as performing a sampled spatial
Fourier transform on the output signals from the array, since they provide a spatial
sampled version of the angular distribution of received signals. This obviously
points to the use of on line computers to perform fast Fourier transforms (FFT) to
form multiple beams. Digital FFTs are being used for Doppler spectrum analysis of
radar signals and for both Doppler and bearing analysis for sonar signals. The power
of such processors is increasing, particularly due to the development of dedicated
computer systems, and their performance is now adequate for application to
medium resolution radar systems. FFTs are now a most suitable choice for multiple
beam forming for either CW radars or low resolution pulse radars. Digital FFT
analysers have already been employed for airborne and spaceborne synthetic aper-
ture radars.9

v array

loca I oscillator

bipoI a r
A/ D A/ D video signals
convertor convertor

F.F.T. processing

output

Fig. 13.3 Formation of multiple receiving beams by fast Fourier transform (FFT) processing
of bipolar video signals from receiving array

The use of such digital processing in general involves translating the signals down
to a low IF which in many cases may be a bipolar video (zero frequency IF). The
signals are then sampled and fed via an A/D convertor into a digital FFT processor.
Such an arrangement is indicated in Fig. 13.3. It is interesting to note that the sche-
matic representation of a FFT process is identical to the physical representation of
the Butler matrix network which was developed several years earlier. There are a
number of processing techniques capable of providing either the Fourier transform
or the angular power spectrum of signals received by an antenna array. Many of
336 Array signal processing
these schemes may also be realised in terms of either surface acoustic wave (SAW)
or charge coupled device (CCD) technologies.10 Recent research is also looking at
optical signal processing techniques for beam forming.11

13.2.2 Multiple beams from element sampling


An alternative way of forming a set of multiple beams from a linear array is to
sequentially sample the output of each element of an array at a repetition rate oos =
1/r and present the resulting signal to a bank of filters having centre frequencies at:
co0, ), (co 0 -

where co0 is the carrier frequency. The output from each filter will have the direc-
tional characteristics of a single fixed beam in space and the set of outputs will cor-
respond to a family of overlapping beams whose positions, beamwidths and sidelobe
levels can be controlled by the choice of filter characteristics.12 A schematic dia-
gram of such a system is shown in Fig. 13.4. In order to ensure that the noise per-
formance is not worsened by the wide bandwidth sampling switch, it is necessary to
define the noise bandwidth of the system by narrow band filters prior to the
sampling. It is possible to employ a stage of amplification at RF or alternatively to
frequency change and incorporate the sampling and the amplification at IF.

beam filters switch


period -1 IT

amplifier
BW -nlf
array

centre
per filter frequency u>0

Fig. 13.4 Formation of multiple receiving beams by Nyquist rate time sampling of the out-
put from elements of a linear array and subsequent band-pass filtering

The equivalent antenna gain of the system would be the same as for other forms
of multiple beam receiving system of comparable front end noise factor. It is worth
emphasising that the full receiving antenna gain applies, despite the fact that only
one element at a time is being used.
When used as a radar system the sampling rate of the element outputs must cor-
respond at least to the radar bandwidth. Therefore for a pulse duration of r the
bandwidth in the medium is 1/r and after the sampling switch is at least n/r (for an
H-element array). If n orthogonal beams are formed by the filtering process (i.e. a
set of rectangular filter characteristics) then this gives n output channels each of
Array signal processing 337

bandwidth 1/r as for other multiple beam configurations. Filtering can be designed
to provide non orthogonal beams if required and digital filtering may also be used
at this stage.
Although the above concept has been tried only experimentally for a radar appli-
cation13 it is in general use in the navigational aid field but used on transmission.
The Doppler VOR beacons involve switching a transmitting source sequentially
around the elements of a circular array in order to radiate a signal which is effec-
tively frequency coded in space. The more recent developments of Doppler micro-
wave landing systems employ the same techniques on linear arrays.4 Here the signal
received in space from the sampled transmitting array is characterised by a
frequency shift proportional to the angle off boresight. The latter development also
incorporates a range of advanced signal processing schemes to control beam shapes
and minimise errors due to multipath effects.

13.2.3 Equivalence between multiple beams and Nyquist rate beam scanning
A direct alternative to multiple receiving beams is to employ a single receiving beam
and to scan this across the sector of interest, repetitively, at the Nyquist sampling
rate. This is equivalent to using a single receiving beam to sample the fixed beam
positions of a multiple beam receiver. We can again refer to our sector coverage
radar system to describe this concept.

sector illuminated
wide-beam
transmission

single receiving
beam scans across
sector at Nyquist rate

Fig. 13.5 Sector illuminated by broad-beam transmission employing a single receiving beam
scanning across this sector at the Nyquist scanning rate

Fig. 13.5 shows a sector illuminated with a floodlight transmission with a pulse
of duration r. A narrow receiving beam is scanned repetitively across this sector at a
rate corresponding to 1/T. The scanning rate therefore corresponds to the Shannon
sampling rate or Nyquist rate. This means that pulses reflected from each target will
be sampled by the scanning receiving beam. The technique is in operational use in
338 Array signal processing

the sonar field,3 but the radar application has been mainly restricted to experimental
equipments.14'15j 16
There are several alternative system configurations for providing Nyquist rate
scanning of a receiving beam and the most well known is shown in Fig. 13.6. Here,
the output of each array element is subject to single sideband frequency translation
by angular frequencies of:

(where to s , the beam scanning rate, = 2IT/T). Such a scheme effectively adds phase
shift at uniform rates to each antenna element, thus producing a uniform move-
ment of the array directional pattern along the sin# scale. Each diffraction maxi-
mum of this pattern represents a successive beam which scans across the sector.

7 receiving array
I single sideband
| / j ~ l modulators
(n-1)u>s
amplifiers (bandwidth MY)

combined output
(bandwidth n / r )

Fig. 13.6 Modulation scanning system to produce repetitive scanning of receiving beam at
rate of u>s
u?o = 2-nlr

The noise factor of such a receiver is identical to that of a corresponding conven-


tional receiving array provided that the bandwidth of the filters (or amplifiers) in
each channel is restricted to 1/r. This bandwidth restriction has been referred to as
a storage filter in some papers.15 Although the bandwidth in each channel is 1/r, the
total bandwidth after summation of the channel outputs becomes n/r, correspond-
ing to that of the sampled pulses (duration rjn). Thus the single output represents
the signals from n separate directions in space which corresponds to the w-fold
increase in bandwidth. The maximum range performance is not affected by the
sampling process and corresponds to the previous examples of floodlight transmis-
sion discussed above.
The main technical problem with the above configuration is achieving single side-
band modulation with adequate rejection of unwanted modulation products. The
use of complex filters in each channel is not very attractive since it is more difficult
to maintain very good phase accuracy through multi-stage filters. In some cases, it
may be easier to incorporate the local oscillator in the single sideband modulation
stage. In this case, simple modulators may be used, with a set of local oscillator
frequencies given by:
Array signal processing 339

co,, (o)i + co s ), (GJZ 4- 2co s ),. . . , (coj + n - 1 (o 8 )


The generation of such phase locked signals will in general require a frequency syn-
thesiser.

n-element array

RF a m p l i f i e r and filter
( b a n d w i d t h UDS)

single-sideband
modulators

spectrum and time waveform at A

spectrum and time waveform at B

post- spectrum and time waveform at C


detector
| filter to
taper
sidelobes
output D
spectrum and time waveform at D

Fig. 13.7 Frequency filtering to apply low sidelobe amplitude tapers to the output of a
receiving array incorporating within pulse modulation scanning

The range resolution is still determined by the transmitted pulse duration r, and
the sidelobe performance is set by the one-way sidelobes of the receiving array. The
scanning receiver has the useful feature that array amplitude tapers may be achieved
340 Array signal processing

by means of filters situated at the output of the combined array. This is illustrated
by Fig. 13.7. Here, the output of a receiving array is subjected to RF amplification
and frequency translation by integral multiples of the beam scanning rate cos. The
signals are then combined and translated to IF. If we consider the IF time waveform
(and its associated spectrum) due to the excitation of the array by a single far field
source, the envelope of the time waveform will be a sin npj'sinp function and repre-
sents the directional pattern of the receiving array. This repeats with a period
2IT/(JOS. This is shown in Fig. 13.7 (output A). The spectrum of this waveform con-
sists of n uniform amplitude spectral lines, one line for each element (or receiver
channel) of the array.
It is evident that if this spectrum is subjected to an IF filter having a character-
istic Tt(co) it acts as a taper function Ti(x) applied to the receiving array. This is
obviously a simple way to add or change a taper, but the laws of circuit theory
present certain constraints on simultaneous control of amplitude and phase, which,
for example, preclude the use of discontinuous taper functions. Such constraints do
not occur for direct control of the phase and amplitude of the signals from the ele-
ments. This situation is depicted at B in Fig. 13.7. These are also examples of post
amplifier tapers.
A further modification of this concept is indicated at C in Fig. 13.7. Here, the
video (post detected) spectrum and wide bandwidth waveform of the scanning
beam is shown. For the simple case of square law detection and a uniform array,
the post detected spectrum is triangular and the corresponding detected beam shape
is (sin np/sinp)2. It is now possible to apply post demodulation filters to suppress
sidelobes. This will now be a low pass response T2(oS) (Fig. 13.7D). The Fourier
transform (taking account of both positive and negative frequencies) will give the
resultant demodulated directional pattern which incorporates the effect of the
amplitude taper.
This type of scanning system is essentially a continuous process at a fixed scan-
ning rate; it is not possible to stop the beam or vary the beam dwell time. It is how-
ever possible to produce the effect of a fixed beam by means of a sample gate at
one of the outputs A-D of Fig. 13.7. If this gate opens repetitively at a rate us for
a period of time r/n corresponding to the instant when the scanning beam is look-
ing in the required direction, its output will correspond to a fixed beam and its
angular location may be controlled by the timing of the gate. This type of output
may be useful to generate a single beam or a pair of tracking beams.
An alternative configuration for within pulse scanning of a receiving beam is that
shown in Fig. 13.8. Here, the output of the elements are mixed with a frequency
swept local oscillator and the difference (IF) is fed into a tapped delay line. The
output of the delay line represents the receiving beam output and the angular direc-
tion of the beam (on a sin Q scale) is linearly related to the frequency of the local
oscillator. The two ends of the delay line correspond to beams deflected either side
of boresight but in practice, it is only necessary to employ one such output. Some
configurations have located the tapped delay line in the feeds from the swept local
oscillator to provide the same effect.
Array signal processing 341

This concept may also be regarded as 'frequency scanning at IF' and in contrast
with the continuous scanning scheme of Figs. 13.6 and 13.7, is also capable of vari-
able scanning rates and stationary beam positions. Since the typical phase deviation
required for wide angle scanning is of the order 2?rrad between adjacent elements,
this can be achieved by means of either a wide sweep Af or a long time delay in
each section of the delay line tx. This may be expressed as:

Phase deviation = Aooti = 2TT or Af = — (13.1)


t

receiving array
Y Y Y (n sections)
Y Y
bandwidth restriction

crystal mixers fed in


phase from local
oscillator

delay line (n-1 sections)

output to intermediate frequency-swept


frequency circuits local oscillator
sawtooth
modulation

Fig. 13.8 Frequency-swept local-oscillator system for electronic scanning of a receiving beam

For the case of repetitive scanning, it is necessary to use a sawtooth frequency


modulation of the local oscillator and this can involve problems in the time taken
for the 'flyback' to propagate down the delay line. This means that the delay in the
delay line must be kept significantly less than l/cos.

13.2.4 Sidelobe levels in floodlight transmission systems


It has already been pointed out that the use of a floodlight transmitter means that
the overall sidelobe performance will be the one way sidelobe level of the ASP
receiver. This implies a limitation on the maximum dynamic range of target echoing
areas; otherwise large targets will paint on displays due to sidelobe returns. The
receive sidelobe level is more important than the two way sidelobe performance in
terms of the rejection of interfering signals and deliberate jamming. Although the
requirement for low sidelobes places severe constraints on the amplitude and phase
tolerances in the ASP receiver, this can be met by stable components and digital
processing. A further powerful technique that can be applied here is to design
closed loop control of phase and amplitude within the receiver.17

13.2.5 Compatibility of ASP with other radar processing schemes


In assessing various forms of processing, it is always important to establish to what
extent such schemes provide inherent advantages and restrictions on their use in
342 Array signal processing
different applications. In this respect, the ASP schemes discussed above are fully
compatible with all forms of coherent radar processing such as pulse Doppler and
MTI. In some respects, they provide significant advantages in this area since the
basic spectral line widths of coherent signals are not broadened by an ASP receiver
(i.e. scanning modulation is absent).
If a mechanically rotating antenna is used to receive a single frequency wave, the
bandwidth of the signal at the antenna output will be broadened due to the
modulation effect of the rotating antenna pattern. This same effect may also be
considered in the frequency domain, in terms of the Doppler shifts imposed upon
the signals received by parts of the antenna which are moving relative to the centre
of rotation. This spectrum broadening implies a limit on the ability of a coherent
radar to distinguish slow speed targets from fixed clutter. Broadly speaking, target
speeds must be above the circumferential velocity of the tip of the rotating antenna.
Clearly, such effects may be overcome by the adoption of fixed multiple beams.
Conventional electronic scanning systems do not overcome the effect since such
Doppler components are simulated by the rate of change of phase in the antenna
phase shifters. For a phased array system with complete control of the beam
position, it is possible to minimise the effects of scanning modulation loss by step
scanning and by ensuring that the beam dwell time in each direction is adequate to
minimise this effect. Needless to say, this is a restriction of the mutual compata-
bility of electronic scanning and the use of coherent signals for the study of low
speed targets.
It is interesting to note that in the case of an ASP system incorporating very fast
electronic scanning (i.e. Nyquist rate scanning) the effect of modulation scanning
loss is absent. This is because the equivalent Doppler components are now multiples
of the scanning rate.
There is no reason why pulse compression or matched filter processing should
not be incorporated with array signal processing schemes. Indeed, one form of
radar8 makes specific use of this feature. However, it is necessary to point out that
since an ASP scheme can be considered as equivalent to n separate radars, it is not
surprising that this will result in either n separate pulse compression receivers or one
pulse compression receiver having a total time bandwidth product n times larger
than that needed for a corresponding single radar.
This latter point applies to various types of radar processing, i.e. coherent sys-
tems, log amplifiers, CFAR circuits and other related schemes. It is, in general,
necessary to provide one such processing system for the output of each beam of the
multi-beam receiver. Fortunately, the cost of such processing in terms of digital
equipment is continuing to fall, thus making this a less severe penalty.

13.3 ASP in combined transmit and receive antenna systems

The directional pattern of a radar system is the two-way spatial product of the
transmitting and receiving directional patterns. In most systems, this merely means
Array signal processing 343

squaring the one-way pattern of the antenna. There are many examples where there
is a substantial advantage in employing different patterns on transmission and
reception. For example, the use of different transmitting and receiving patterns in
radar arrays can lead to improvements in either beamwidth or sidelobe perform-
ance. The systems discussed in the previous section represent examples where the
use of a floodlight transmit pattern results in a considerable simplification of the
electronic scanning hardware but at the expense of sidelobe performance. We shall
now look at cases where the radar employs different patterns on transmit and
receive to achieve various advantages.

133.1 Pulse compression in angle by within-pulse scanning on transmission


If, instead of illuminating the sector of interest with a single broad beam, floodlight
transmitter, a narrow transmitting beam is switched on for a period t2, and simul-
taneously scanned across the sector for the period t2, then the duration of the trans-
mitted signal illumination in one direction will be given by:

°ft2 (13.2)
Here, 6b = the beamwidth of the transmitter and 6S = angular width of the sector.

Fig. 13.9 Floodlight illumination of a sector by means of a narrow beam scanned across the
sector in a time t2

The effective duration of the transmitter pulse sent out in any given direction by
this system is set by the effective transmitter beam dwell time in that direction. The
system therefore provides a pulse compression in angle effect which is illustrated in
344 Array signal processing

Fig. 13.9. The within pulse scanning system shown in Fig. 13.6 may be reversed for
operation as a transmitter to achieve this effect, but the transmitting system will
need high power transmitting amplifiers at each antenna element.
This angular pulse compression transmitter may be used with any of the ASP
receiving systems discussed earlier and there is no need for either the beamwidth or
scan rates to be the same between the transmitter and the receiver. The scanning
transmitter is here being used only to achieve a pulse compression effect and the
overall directional pattern of the radar is still the one-way response of the receiving
array despite the use of a directional transmitter. This is because we are effectively
achieving an omnidirectional coverage of the sector from a scanned transmitter.
The above simplified argument does not take account of the sidelobes of the
transmitted signal which will now give rise to range sidelobes in the resultant radar
performance.

13.3.2 Array thinning on both transmit and receive by the use of ASP
Fig. 13.10 shows an interesting array configuration employing a short filled array
on transmission and a wide spaced long array on reception. By choosing the length
of the short array to equal the inter-element spacing of the long array, the resulting
two-way spatial pattern corresponds to a sin«x/sinx pattern. It will be seen from
this Figure that the diffraction maxima of the receiving pattern are cancelled by
multiplication with the directional zeros in the transmitting pattern. If there are nx
elements in the short filled array and n2 elements in the long array, the resulting
pattern corresponds to a one-way pattern of an array containing n\.n2 elements.
Clearly, this corresponds to a substantial economy in elements for a given angular
resolution, but we also need to examine the limitations of such a scheme.
First of all, it is evident that the combined T & R sidelobe performance of this
system is equivalent of a one-way sidelobe performance (in the absence of ampli-
tude tapers). Furthermore, the sidelobe performance on receive only will contain
diffraction sidelobes at the OdB level corresponding to the diffraction maxima.
This may not be a severe disadvantage for radar performance if the dynamic range
of target echoing area is not large, but it is obviously a severe problem for any inter-
fering signals or jamming.
The directional gain of the receiving aperture will be set by its physical receiving
aperture (taking account of tapers and the wide interelement spacing). This is not
the gain associated with the directivity of the two-way pattern. At first, it would
appear that a further disadvantage with the scheme is that the power in the broad
beamwidth transmitter lying outside the angular coverage of the narrow receiving
beim is wasted. However, we can now recover all this power by generating a set of
multiple narrow beams from the long wide-spacing array, so that these beams cover
all the sector illuminated by the transmitting array. This corresponds to ASP within
the sector of the transmitted beam. Any of the ASP techniques discussed in the pre-
vious sections could be used for this purpose.
It is also interesting to note that the short array may now be subjected to elec-
tronic scanning on transmission alone (requiring only nx phase shifters). The
Array signal processing 345

receiving array would then employ n2 elements and incorporate array signal proces-
sing within the transmitted section. This still gives the required two-way capability
corresponding to n1n2 elements and the ability to steer the coverage sector over a
wide angle (typically ± 45°). This is a very economical combination, but the limita-
tion on the technique is the sidelobe performance against interfering signals. This is
an area where a null steering would be very valuable to reduce such interfering sig-
nals. The topic of null steering is discussed later in this chapter.

TX pattern

RX pattern

V V wv

RX TX

Fig. 13.10 Use of a short filled transmitting array plus a receiving array with wide inter-
element spacing to provide a 2-way directional pattern having the form (S\nnp/
Sinp)

Although the above transmit and receive ASP system has been discussed for one
dimension of electronic scanning, the technique lends itself very easily to a two-
dimensional electronic scanning system (i.e. three-dimensional radar coverage). Fig.
13.11 shows transmitting and receiving linear arrays located at right angles in space
and the two-way product corresponds to a pencil beam at the intersection of the
two fan beams. It is then possible to scan this pencil beam in both azimuth and ele-
vation by separate phasing of the transmit and receive arrays.
This arrangement may be thought of as a transmit and receive version of the Mills
Cross array22 used in radio astronomy. Again, there will be transmitted power lost
due to radiation into directions of the main beam not covered by a corresponding
346 Array signal processing
receiving beam and again, this energy can be recovered if the receiving signal
employs array signal processing with either multiple beams or Nyquist rate scan-
ning.

'TX array

RX array and BFN

Fig. 13.11 Use of crossed transmitting and receiving beams to provide a 2-way pencil-beam
configuration

, m u l t i p l e receiving
-fan beams

2-way pencil beams

C \fan transmitting beam

Fig. 13.12 Formation of 2-way pencil beams from product of T& R fan beams without loss
of power

Fig 13.12 shows a section through such a set of beams and indicates how the
multiple receiving beams then generate a set of pencil beams which may be steered
in elevation by the phasing of the transmitting array.
Array signal processing 347

133.3 Multiple beam operation on transmit and receive


The multiple beam system discussed in Section 13.2.1 may be extended by employ-
ing an identical set of multiple beams for both transmission and reception. To use
this configuration as an ASP system, it is necessary to employ different transmitted
frequencies in each beam, otherwise the resulting transmitted pattern will still be
that of a simple floodlight system. (It is also possible to use time diversity multi-
plexing of the transmit beams but at the expense of a severe loss of data rate). This,
therefore, corrresponds to n separate radars on n different frequencies operating via
n separate adjacent beams. Clearly, the overall sidelobe performance is the square of
the one way pattern.
The advantage of this type of system over a simple floodlight system is that the
use of multiple transmitters enables a higher power to be radiated and the overall
sidelobe performance of the radar is the two-way performance of the antennas used
on transmit and receive. This type of configuration has been used for a number of
stacked beam 3D radar systems.5 The use of different frequencies on adjacent beams
means that the echoing area of a target will not be the same on adjacent beams.
This limits the capability of amplitude comparison of signals from adjacent over-
lapping beams to achieve improved angular accuracy (within the beamwidth), unless
monopulse receiving channels are provided for all the frequencies employed.
The mechanical problems of rotating a large number of transmitters and
receivers with a large antenna can be fairly complex. Multiple channel waveguide
rotating joints have, however, been developed which are capable of coupling several
high power transmitters to a set of independent multiple beams on a lens or reflec-
tor antenna system. Fig. 13.13 shows a photograph of an experimental multi-beam
radar at RSRE Malvern.
The above system uses multiple transmitting beams to code different directions
in the sector with a different radiated frequency. Two other methods described
below may be used to achieve this effect, but they both result in a frequency vari-
ation which is a continuous function of angle from boresight. In the simplest cases,
this corresponds to a frequency variation proportional to the sine of the angle from
boresight.

13.3.4 Within pulse time sampling of transmitting array


This corresponds to reversing the scheme of Fig. 13.4 to produce a transmitter. The
output of a transmitter is sequentially commutated along the line of antenna ele-
ments. The resulting radiated signal approximates to a sector illumination where the
frequency of the transmitted signal is proportional to sin 6 (the angle of boresight
from the array). This commutated array approximates to a single transmitted
source moving along the array at a constant speed and the different frequencies
radiated in each direction correspond to the Doppler shift of the transmitted fre-
quency due to this movement as seen from each particular direction in the sector.
In practice, the commutated array is a repetitive quantised movement of the source
and this leads to a transmitted line spectrum. The maxima of the envelope of this
spectrum changes with angle which produces the effect of radiating a different
348 Array signal processing

frequency in each direction. This is the basis of the Doppler Microwave landing sys-
tem.

13.3,5 Within pulse frequency scanning


If a frequency scanning antenna is excited by a within pulse frequency swept trans-
mitter, then a single swept pulse results in the radiated beam moving across the sec-
tor, thus coding each direction in space with a different frequency. The frequency
modulation of the transmitter may be used to produce a shaped beam since the
transmitter's spectral density is inversely proportional to the rate of change of
frequency. This technique is used in the Plessey AR3D radar8 which is shown in
schematic form in Fig. 13.14. Here the ASP scanning system provides the vertical
coverage and is mechanically rotated to give 3D information.

• t * *m*

Fig. 13.13 Radar antenna with linear array feed to generate stacked beams in elevation

The shaped modulation of the transmitters frequency sweep produces an ampli-


tude spectrum, which to a first approximation may be written as:

If the dispersive characteristics of the frequency scanning feed is G(0,/) the


Array signal processing 349

resultant directional pattern comprising the vertical coverage is given by:

= (13-4)

The received spectrum is the product of the transmitted spectrum with the square
of the antenna spectrum at the particular angle of interest. The resultant receiving
beams may therefore be defined by a bank of filters (in the case of AR3D by SAW
filters at IF). This radar system provides an overall 2-way sidelobe performance plus
the ability of measuring range and angle of all targets within the sector from a single
transmitted pulse. In addition, pulse compression is applied to all received signals
and the antenna spectrum helps to provide frequency weighting which reduces the
level of the time sidelobes of the compressed pulses.

dispersive
feed

non-linear all-beam-pulse-
•f0c9
FM transmitter compression equaliser

Fig. 13.14 Basic principles of within pulse frequency scanning radar as employed in Plessey
AR3D radar

13.3.6 Directionally decorrelated transmitter beams


The above schemes have used variation of transmitted frequency with angle to code
different directions in space. This is just one method to ensure that signals transmit-
ted in different directions are decorrelated.18 The use of decorrelated transmissions
has two major advantages for ASP systems:
(a) the overall sidelobe performance is a genuine 2-way sidelobe performance (as
mentioned above)
(b) it is possible to use the technique to reduce the effect of interference or deliber-
ate jamming.
With directionally decorrelated transmissions, the average power distribution across
the sector remains uniform, but the transmitted signals differ in the various beam
directions.
Directionally decorrelated beams may be obtained by connecting decorrelated
transmitters to either:
(a) separate beam ports (Figs. 13.15 and 13.16) or
(b) separate array elements (Fig. 13.17).
Receiver processors for method (a) are relatively straightforward since each
350 Array signal processing

decorrelated
transmitters

Fig. 13.15 Directionally decorrelated transmitter beams: separate transmitters per feed horn

beam-forming network

/ \ decorrelated
V /transmitters

Fig. 13.16 Directionally decorrelated transmitter beams: separate transmitters per beam port

Fig. 13.17 Directionally decorrelated transmitter beams: separate transmitters per array
element
Array signal processing 351
transmitted waveform is directly associated with a unique beam direction. With
method (b), the transmitted 'code' in a particular direction is formed from the
uniquely phased sum of the element excitations corresponding to that direction. It
is thus necessary to reconstruct this phased sum for the receiver processor: a general
method of doing this is illustrated in Fig. 13.18. The advantages of method (b) are
that it can avoid the need for a high power multiple beam network.

transmitting array receiving array

correlation
detectors

beam 1

Fig. 13.18 Schematic transmit/receive system for directionally-decorrelated transmitted


beams

13.3.6.1 Separate transmitters per beam port: Basic methods of achieving trans-
mitted beam decorrelation are to employ either:
(a) frequency separation of transmitters or
(b) code separation of transmitters operating at the same carrier frequency.
The first method is relatively easy to implement, but since the available frequency
allocation must be shared between the beam ports, the transmission in each beam
on each pulse must be relatively narrow band. (A set of frequencies for each beam
could be obtained by providing more transmitters than beam ports.)
The second method uses the whole of the frequency allocation for each beam
and is thus capable of much finer range resolution. However, the codes chosen must
be very long if they are to achieve adequate decorrelation. For example, if random
phase codes of m bits duration are employed to simulate noise-type transmissions,
and received signals are recovered by tapped-delay line matched filters, the near-in
time sidelobes in the latter will have an expected amplitude of l/y/m relative to the
peak (the peak values of the time sidelobes may be some 10 dB greater). Thus m
must have a value of 104 to achieve 30 dB time sidelobes. In addition, each receiver
channel will also respond to other transmitted beam codes which illuminate the
target via spatial sidelboes: the level of this unwanted signal will be further attenu-
ated by the cross-correlation process, but will become significant if the number of
transmitted beams exceeds the average spatial sidelobe power ratio.
352 Array signal processing

13.3.6.2 Separate transmitters per array element: In this case, transmitted beam
decorrelation can be obtained by:
(a) frequency separation of element excitations
(b) time separation of element excitations or
(c) code separation of element excitations.
Frequency separation using incoherent transmitters produces noise-like waveforms
in each beam direction. The correlation detectors in the receiver (Fig. 13.18) must
store the amplitude and phase record corresponding to each beam direction. This
implies considerable storage capacity. A more economical method is to employ co-
herent frequency separations for the various transmitter modules, as would be
required to produce a pulse compression in angle-type transmitted scan, but with a
pseudo-random arrangement of frequencies along the array.
radiators

n I I

UUQUQU
phase I random
shifters 1 phase-code
generator

random
pulse firing
modulators I sequence
generator

oscillators
low-power
locking
signal
Fig. 13.19 Phase-coded time-division-multiplex transmissions to produce directionally de-
correlated transmitted signals

Time separation of element excitations is illustrated in Fig. 13.19. Each array


element is pulsed in random sequence, and with random phase. (Since the phase
difference between adjacent elements of an ^-element array required to produce
one beamwidth shift is 2n/n, it is appropriate to use this as the minimum phase
increment.) Each transmitted beam is essentially of constant power with a multi-
level random PSK type of digital pulse compression code. Time sidelobes in the
recovered pulse vary in amplitude from about lj\/n close to the peak to Ijn at the
extremities.
These can be improved by coherently integrating a succession of pulses, if the
coding is changed from pulse to pulse. Code separation of element excitation pro-
duces a good approximation to an entirely noise-like set of decorrelated trans-
mitted beams. As with frequency separation using incoherent transmitters, both
amplitude and phase of the transmitted code corresponding to each beam direc-
tion must be stored for use in the receiver's correlators.
Array signal processing 353
13.3.7 Alternative configurations including multistatic systems
The previous sections have reviewed a wide range of ASP techniques for both trans-
mission and reception. Clearly there is an even larger number of transmit/receive
combinations to handle a given high data-rate for radar or sonar application. It is,
however, evident that the power of ASP processing is most relevant to 3D coverage
problems where the data rate requirements may preclude the use of either mechan-
ical movement of antennas or even the use of full beam agility with phased arrays.
It is important to point out that conventional electronic scanning (i.e. common
T & R electronic beam positioning of a single beam) has no advantage in data rate
over conventional mechanical scanning for a purely surveillance role and in many
cases only a marginal increase for a combined surveillance and tracking role. This
suggests that the principal radar applications of ASP will continue to be in the field
of 3D radar systems where the ASP role is to increase the data renewal rate and pro-
vide one extra dimension of resolution (usually elevation).

13.3.7.1 Multistatic systems: Previous discussions of 2-way directional patterns


have assumed that the transmitters and receiving antennas are co-located, or suf-
ficiently near that the resultant 2-way pattern may be calculated as the product of
the transmitting and receiving patterns. There are a number of cases where there
may be some advantage in departing from this situation.
A simple example is a CW radar where separate T & R apertures are employed to
reduce transmitter leakage into the receiver. If the two antennas are separated by
several times the antenna dimensions, then the resultant beam patterns will not
overlap until the far field region. This gives a 2-way directional pattern which is the
normal 2-way beam response in the far field but with a low '2-way sidelobe only'
response in the near field. This can be of value in reducing unwanted responses
from short range clutter.
If we generalise this situation we get a multistatic radar configuration in which a
combination of transmitters and receivers is distributed among a number of sites
with overlapping cover. This provides additional location information on targets by
combining conventional radar operation (from a common T&R site) with
additional receiving sites to provide additional triangulation data. In this simple
example the location of the target is determined by the target bearing and timing of
the received signal at a number of sites so that the normal definitions of range reso-
lution and angular resolution become interrelated and result in combining a number
of different shaped resolution cells. Such multistatic configurations are currently
receiving significant attention for certain radar applications, where the additional
data may be used to resolve ambiguities due to high prf transmissions and to
observe targets in the presence of directional interference and jamming.
Multistatic configurations usually require the receiving only sites to employ mul-
tiple beam antennas or some other form of ASP in order to simultaneously observe
returns from within a range sector of the transmitting beam. The various techniques
described in Section 13.2 would apply to such applications. Reference 19 gives a
review of Bistatic and Multistatic radar systems and their antenna implications.
354 Array signal processing

13.3,8 Signal processing technologies


Since the directional properties of an ASP receiver are dependent upon the per-
formance of the assocated signal processing system, the potential capabilities of
ASP are currently increasing due to improvements in processing technologies. On
the transmitting side, the required signal processing is the same for both ASP and
conventional electronic scanning (i.e. ferrite and diode phase shifters, high power
transmitters and high power antenna systems such as frequency scanners and Butler
matrix networks).
The growing relevance of ASP relates to continuing developments in receiver
technologies; these include solid state sources, microstrip circuits, surface acoustic
wave delay lines and filters, charge coupled devices and digital integrated circuits.
All these technologies are offering increased performance at reduced size, power
consumption and cost. This is also being achieved with increased reliability and sta-
bility. Since ASP systems in general require a large number of parallel processing
channels it is also necessary to integrate significant proportions of the processing
system into a few complex units. Requirements for low sidelobes can place severe
constraints on such units in terms of gain and phase stabilities over wide dynamic
ranges.
The dynamic range of radar signals varies with different applications and one of
the most difficult cases is that of a ground radar where energy directed near the
horizon may strike large buildings and other sources of point clutter. For such a
radar with a signal/noise ratio of 20 dB for a i m 2 target, the likely upper limit on
the level of such point clutter is in the region of 105m2. This suggests a receiving
dynamic range of the order of 70 dB in addition to the R4 law, which may be diffi-
cult to fulfil.
Such high level clutter will break through the MTI system and can also lead to
unwanted sidelobe responses and intermodulation effects in the receiver. For exam-
ple, the signal received by a radar from a source of clutter of the above magnitude
at lkm range would be comparable in amplitude with a local oscillator signal
applied to the mixer. The same situation would apply to the echo from a pigeon at
10 m or the direct transmissions from a second similar radar at about 30 km. Such
high level signals can therefore give rise to severe intermodulation effects in the first
mixer of the radar receiver.
If a single antenna associated with a single receiver is replaced by a multiple ele-
ment array incorporating separate receivers for each element, the effect of such
high level signals is substantially reduced, due to the smaller gain of each individual
element. This advantage is somewhat offset by the fact that the signals received at
each element will include any high level clutter or interference from the broad
angular directional pattern of the element. If, on the other hand, the receivers are
incorporated after a passive multiple beam forming system, intermodulation per-
formance of each receiver would be similar to a conventional directional antenna/
single receiver system.

133.8.1 Microwave systems and components: Array signal processing receiving


Array signal processing 355

systems can incorporate printed circuit antenna arrays. For ASP systems where the
signal frequency channels are at microwave frequencies, most of the microwave
components such as hybrids, circulators and phase shifters can be constructed using
printed circuit technology, which is now in widespread use. Multiple beam matrix
networks such as the Butler matrix or the Maxson matrix can be constructed at
microwave frequencies in stripline and some Butler matrix systems are commer-
cially available in this form. The absence of high power requirements can make
these components comparatively cheap. An alternative approach is the lens fed
array where a lens, such as the Rotman lens can be used to provide a multiple beam
capability.18 These lens systems may be designed in waveguide or stripline con-
figurations.
The continuing improvements in the performance of solid state microwave
amplifiers is encouraging and it is therefore possible to incorporate limited amounts
of microwave amplification in each channel with good phase and gain stability. This
reduces the need for specialised phase control loops for all but the most stringent
sidelobe requirements.

13.3.8.2 Intermediate frequency processing: At intermediate frequencies solid


state amplifiers and integrated circuit techniques already provide cheap, stable,
analogue sub-systems with good noise factors and moderate dynamic ranges.
Developments of surface acoustic wave (SAW) filters, delay lines and correlators are
particularly appropriate at IF. Such SAW components are most conveniently imple-
mented at frequencies below about 300 MHz, and are appropriate where large num-
bers of filters or delay lines are required to be made from a single mask. This is one
of the reasons why the additional cost of multiple channel receivers in an ASP sys-
tem is not substantial in a complex radar.
The adoption of pulse compression for an ASP radar generally requires one pulse
compression filter per channel and in most cases the filter designs are identical in
each channel. The band limiting filters required for Nyquist rate scanning as well as
pulse compression filters also appear appropriate for SAW techniques. The tem-
perature characteristics of SAW filters can be sufficiently stable to meet the require-
ments of amplitude and phase stability of the reciever channels.
Although charge-coupled devices (CCDs) are usually thought of as a base-band
processing technique, improvements in sampling rate make these increasingly com-
petitive at IF. This technology also looks attractive for producing delay lines, filters
and beam forming systems. The ability to operate at more than a single clock rate
also offers the possibility of convenient changes in system parameters which would
be very difficult with other analogue techniques. A possible limitation of the tech-
nique is that of dynamic range.
Multiple beam forming networks may be constructed at IF using resistors,
hybrid transformers or tapped delay lines. The latter can be produced from lumped
circuits, transmission lines or acoustic delay lines using bulk or surface waves. Beam
forming at IF may be more convenient in many ways since it may be performed
after amplification so that the losses in the beam forming network are less
356 Array signal processing

important. It is therefore acceptable to design networks for non-orthogonal beam


positions after amplification.

133.8.3 Digital and base band beam forming: The concept of digital beam
forming first grew up in sonar systems though it had been used to a limited extent
in radio astronomy. Both these applications are associated with fairly low RF band-
widths, but modern developments in wide band digital processing have extended
the capability to radar. Reference 20 describes the use of digital processing in sonar
systems and Reference 21 is an example of digital beam forming for radar.
The basic concept of digital beam forming first requires frequency changing the
received signal down to a low IF. In general, this is a zero IF (though modified by
Doppler shifts) and the signal is then fed to analogue-to-digital convertors to pro-
duce two streams of binary baseband signals representing 'in-phase' (I) and 'quadra-
ture' (Q) channels. These sets of digital signals then represent the amplitude and
phase of signals received on each element of the array. The process of beam form-
ing implies weighting these digital signals with the appropriate complex values of
amplitude and phase and summing the resultant to form a beam. The resultant sig-
nal may, of course, be converted back via a digital-to-analogue convertor to provide
a beam output. Note that this process is linear and does not involve demodulation
of the signals which still retain phase information.
There are many possible configurations for performing the digital processing. In
the simplest case, dedicated multipliers are used for each channel to multiply the
element signals by the appropriate beam steering weightings. These instructions can
be driven by software and can produce scanning beams, multiple beams or null
steering, as will be discussed later in the chapter.
It is interesting to note that the flow chart of a discrete fast Fourier transform
programme is exactly the same format as the Butler matrix network which evolved
several years before fast Fourier transform programmes became available. The
digital FFT generates a set of orthogonal beams giving the usual —13 dB sidelobes
from an unweighted aperture. Dedicated FFT machines are becoming available for
various functions and can therefore be applied directly to beam forming at base
band. Since digital processing is of necessity performed after amplification of the
signals, the relevance of orthogonal beams is far less important.
Many different configurations are thus available for scanning beams or generat-
ing multiple beams and this can be done with dedicated special purpose hardware
or in general purpose computers, though the latter are rarely particularly efficient
for this purpose. Sampling rates of about 20 MHz are currently available though this
does tend to limit this form of processing to low or medium range resolution radars.
The number of bits employed in the processing systems and the design of the soft-
ware also places limitations on the dynamic range of the signals and on the resulting
sidelobe levels of the beams so formed. Although the concept has been described
here for reception, it is clearly possible to use digital processing to generate suitable
IF signals which could represent array excitations for any beam direction. They
could thus be amplified for electronic scanning on transmission if required.
Array signal processing 357

13.4 Multiplicative signal processing and post demodulation synthetic aperture


techniques

Multiplicative signal processing has been used extensively in radio astronomy but
has had only limited application in radar. In radio astronomy it has been used for
three different types of application:
(a) To achieve 2-element interferometer arrays.
(b) To synthesise the pencil beam of a square aperture by multiplying the output of
two crossed linear arrays22 (a special form of interferometer called a Mills Cross
array).
(c) For aperture synthesis by post demodulation processing;2 here the product of
signals from pairs of elements located at different positions on the ground can be
added together to synthesise the output corresponding to a large filled aperture.

In both radar and sonar, multiplicative arrays have been used to achieve small
reductions in beamwidth for various receiving arrays. This technique had also been
adopted in one form of harbour radar.23 A feature of multiplicative arrays with
potential for exploitation in radar systems is the ability to produce normal beam-
width patterns from thinned arrays containing very few antenna elements.

13.4.1 Interferometers
One of the early uses of 2-element interferometer arrays was in radio astronomy
where the need for increased angular accuracy and reduced beamwidth requires
antennas having very large dimensions. This problem was solved by the use of two
antennas having a large spacing between them. If the outputs of two such identical
antennas are combined and fed to a receiver incorporating a detector, the resulting
direction pattern is given by:

D{6) = De(6)cos-sin6 = \De(0)W)\ (13-5)


X
where De(6) is the directional pattern of the element of the 2-element array and / is
the spacing between the two antennas. In the above equation, 1(6) is the interfero-
meter pattern of a 2-element array. A modulus sign is included in the expression
since this section will be discussing alternative forms of demodulating the output of
an antenna array and it is important to distinguish between the output of a direc-
tional receiver incorporating some form of demodulation signal processing and the
normal field variation.
Fig. 13.20 shows an interferometer array and its directional pattern for the case
of ideal modulus detection and for square-law detection. This could represent a
typical output of an interferometer array when this pattern is scanned across a far
field source such as a radio star. Clearly, the angular accuracy of the co-ordinates of
this source can be substantially increased by the use of the interferometer array due
to the reduced lobe width of the interferometer term compared with the lobe width
of the pattern of a single element. This accuracy can be further improved by
358 Array signal processing
increasing the spacing between the two elements. This will be eventually limited by
the ability to determine which interferometer lobe is the central lobe of the result-
ing pattern when the signal is contaminated by noise. For radio-astronomy applica-
tions, the principal interest is usually in the power of the received signals, so it is
convenient to employ a square law detector giving an output directly proportional
to incident power. For this case:
1 2irl
D(B) = ) - 1 + cos sinfl (13.6)
2 X

•De(9)

|D e (9)|

e(6) 1(0) |

Fig. 13.20 Two-element interferometer array and its directional pattern with modulus detec-
tion

The above concept depends upon the fact that there is only one source within the
broad beamwidth of the element patterns, otherwise the interferometer response
will be a complex signal which is difficult to interpret.
When any directional receiver is scanned across a far field distribution of signal
sources 5(0), the output of the directional receiver is the convolution of the direc-
tional pattern of the antenna with the angular distribution function 5(0).

D(6)*S(0) (13.7)
This situation applies for either single or extended sources and in the latter case can
be used to measure the angular width of the radio source.
Array signal processing 359

13.4.2 Interferometers with multiplicative processing


A modification of the above interferometer system is to replace the square law
detector with a signal multiplier as shown in Fig. 13.21. Here, the outputs of the
two directional antennas are fed to a circuit which provides an output signal pro-
portional to the product of the two inputs. If the two received signals take the form
of cos cjt and cos (cor 4- \jj) (where \Jj = 2irl sin 0/A), the output of the multiplica-
tion will be:
D(6) = (13.8)

Fig. 13.21 Two-element multiplicative interferometer array and its multiplicative (bipolar)
directional pattern

The 2co term may be filtered out leaving the demodulated signal having a directional
pattern of the form:

D(d) = De(8) cos I — sin 0 = De(6)Im(p) (13.9)


360 Array signal processing
where / m (0) is the interferometer term for multiplicative processing. A multiplier
has a square law characteristic so that these output signals are again proportional to
the power of the incident signals. It can also be seen from Fig. 13.21 that these
directional patterns are bipolar functions, having both positive and negative polarity
lobes. The multiplicative response also contains twice as many zeros as the
corresponding response of Fig. 13.20. Although it is common practice to draw con-
ventional directional patterns having both positive and negative sidelobes, these
effects are not observed in practice due to the use of rectifying forms of detectors.
It is interesting to note that both square law interferometers and multiplicative
interferometers contain the same basic directional information but the square law
system also contain unwanted information corresponding to the rectified output of
the individual elements alone. For example, we may write the directional response
of a simple square law detected interferometer array (taking account of the detector
law) in terms of the signals received by the two elements (i.e. ex and e2).
Square law detected output = e\ + e\ + 2^^2
Clearly, for the case of omnidirectional elements, the first two terms in the above
expression do not contribute to the high resolution angular information, and when
they are omitted, we obtain the same demodulated signals as a 2-element multipli-
cative interferometer array. These element self product terms do contribute to
the detection of a source (or target) but not to its angular location.
There are various circuits that can be used to multiply signals together including
the 'quarter squares multiplier' employed in analogue computers. Radio astronomy
antennas often use the phase switching form of multiplier where the outputs of the
two antennas are combined together in a circuit which periodically reverses the
phase of one input at a convenient switching frequency. The output then contains a
modulation product centered at this frequency whose amplitude represents the
product of the two inputs. Current technology suggests the use of digital techniques
to perform such multiplication by translating the outputs from the antennas to a
convenient low IF. It is obviously necessary to ensure that the multiplier can handle
the very wide dynamic range of the output signals from multipliers.

13.43 Principles of multiplicative arrays


If two separate antennas A and B (see Fig. 13.22) have directional responses given
by Da(6) and Db(6), then the output of the multiplier will have a demodulated
directional pattern Dm(6), given by:
Dm(0) = Da(d)Db(d)Im(d) (13.10)
where Im(6) is a 2-element multiplicative interferometer term due to the separation
of the phased centres of these two antennas. By suitably choosing the form of the
directional pattern, it is possible to achieve either a reduction in beamwidth24 (by a
factor of about 2) or a thinned array system.25
A word of caution is, however, necessary in considering such schemes, since the
output of the multiplying process possesses a square law characteristic. When this
Array signal processing 361

typical multiplicative
array configuration

wide-aperture
response
Da(6)

short-aperture
response
D b (6)

overall multiplicative
directional pattern

interferometer
function

Fig. 13.22. Composite directional patterns of a general multiplicative array configuration

F ig. 13.23 Three-directional responses of an 8-element linear array


Additive linear rectified
Additive square-law rectified
Multiplicative 4 x 4
362 Array signal processing
characteristic is taken into account, it is often found that multiplicative arrays pro-
vide an unacceptably high sidelobe level.
For example, Fig. 13.23 shows three different forms of directional pattern of an
8-element linear array. The first case corresponds to modulus detection of the
additive output; we then have:
sin8p
D(0) = (13.11)
8 sin/?
(where p = (nd/X)sind and d = inter element spacing). The second case refers to
square law detection; the above expression then becomes squared. The third case is
an example of multiplicative processing (where the signals from the four end
elements are added together and multiplied by the signals from the other four); we
then have

It can be seen from Fig. 13.23 that the multiplicative pattern produces a reduced
beamwidth but a higher sidelobe level than the additive pattern. However, we have
not taken into account the square law response of the characteristic. Clearly there is
no directional advantage in employing square law detection compared with
modulus detection; indeed, if the first two characteristics were plotted on a true dB
scale (in terms of the variation of the input signal with direction necessary to main-
tain a constant output from the array) they would produce identical responses.
This method of plotting directional responses in dB may therefore be adopted in
order to remove the effect of the detector law (or demodulation law) on the result-
ant directional patterns. Under these conditions the first sidelobe level of the two
additive patterns will be — 13dB, but that of the multiplicative pattern will be
about —7 dB.
When a pattern is subjected to square law detection the number of zeros in the
demodulated response is effectively doubled, due to each zero becoming a double
coincident zero. Multiplicative processing also doubles the number of zeros in the
resultant pattern but they need no longer be coincident zeros and this gives some
increase of flexibility in synthesis of multiplicative patterns. In general, the use of
multiplicative processing only produces reductions of beamwidth at the expense of
some increase of sidelobe level when the effect of the square law response is taken
into account. Claims for improved resolution due to multiplicative processing
should also be viewed in terms of the overall effect of the processing on sidelobes.
Some of the literature has tended to regard multiplicative arrays as fundamen-
tally different from conventional arrays. However, if we make comparisons between
demodulated additive arrays and multiplicative schemes, there is no fundamental
difference. Both systems produce various combinations of self-products and cross
products of signals from the elements.
The problem of pattern synthesis can be handled by applying conventional syn-
thesis techniques to the two sub arrays. The resultant pattern is then the product
Array signal processing 363

of these two patterns and the interferometer directional term. One of the most con-
venient ways of designing any demodulated directional pattern is to analyse it in
terms of spatial spectra.

13.4.4 Spatial frequency responses for multiplicative arrays


The previous sections drew attention to the fact that the multiplicative output of
two antenna elements is very similar to the corresponding output when the signals
are added and square law detected. The former contains only a cross product term,
whereas the detected output also contains self product terms for each element. This
has led to the use of the concept of a spatial frequency spectrum to analyse and
represent any multiplicative array combination. The spatial frequency response of
an array is the Fourier transform of its demodulated directional pattern. The con-
cept therefore takes full account of the form of signal processing employed and can
therefore compare different forms of demodulation as well as different array con-
figurations. The spatial frequency spectrum offers greater insight into the operation
of various forms of directional systems.
Consider an additive array in which the output signals are summed and then sub-
jected to square law detection. The square law process corresponds to an auto-
convolution of the aperture excitation which therefore yields the corresponding
spatial frequency spectrum. For a multiplicative array, the spatial spectrum is given
by the cross correlation of the aperture distributions of the two consitituent arrays
forming the multiplicative pair. For example, Fig. 13.24 shows the directional
pattern and spatial spectrum of a square law detected array. This takes a triangular
shape but it is sometimes drawn in a symmetrical form to show the corresponding
presence of negative spatial frequencies. For some types of arrays it is important to
show both components of the spatial spectrum.

sinxi 2 detected
x J nresponse

0
spatial frequency spectrum

imimiiimiiiii
square-law
detector

Fig. 13.24 Square-law detected linear array and its directional and spatial frequency respon-
ses
364 Array signal processing

Each component in this spectrum may be considered as an indication of the res-


ponse due to a particular element separation. For example, in an ^-element array
there are n terms with a zero element spacing (the self product terms). There are
(n — 1) terms with a spacing d and (n — 2) terms with a spacing 2d. This gives a
physical picture of the above spectrum.
This can be extended to multiplicative arrays. Fig. 13.25 shows the directional
pattern and spatial spectrum of a multiplicative array consisting of a short filled
array of five elements and a co-centred thinned array also consisting of five
elements. This arrangement gives a uniform spatial frequency spectrum in which
each spectral component comprises two component terms. In Fig. 13.25 the result-
ant multiplicative directional pattern will correspond to the product of the con-
tinuous pattern and the dotted pattern. This product will correspond to the pattern
of a filled array (notionally containing 25 elements). This is an example of using
multiplicative processing to produce a thinned array. There are two disadvantages
with such schemes. The first is that the normal directional pattern produced by
such a uniform spectrum corresponds to a sin me/sin x multiplicative pattern. This
function has —13 dB sidelobes on a square law scale but this corresponds to real
sidelobes of 6.5 dB when considered in terms of the ability to reject sidelobe signals.
This type of thinned array has been analysed by Ward26 who has shown that it is
possible to synthesise tapered distributions in order to simultaneously achieve a
high degree of thinning plus low sidelobes. Fig. 13.26 is plotted on a true dB scale

IT ? T 1 MM ?{• M

spatial frequency spectrum

Fig. 13.25 A 5 x 5 thinned multiplicative array with the constituent directional pattern and
spatial frequency spectrums

as discussed in Section 13.4.3 and shows a theoretical low sidelobe pattern which
simulates the response of such a thinned array using only twenty eight elements to
synthesise a seventy four element array.
The second disadvantage of the above scheme is the complex response of multi-
plicative arrays when illuminated by multiple sources. This will be discussed in
more detail in the next section.
Array signal processing 365

The spatial frequency concept shows how the response of an array is made up
from constituent components each related to the inter-element spacing of pairs of
elements. Since the output from such an element pair spaced say d apart corresponds
to the same output, irrespective of where the pair occurs within the array, we can
argue that each element spacing should not be duplicated as this corresponds to a
waste of elements. In terms of the single source response of an array this is quite
true. An array which does not duplicate its inter-element spacing (or which has only
one spatial component for each spatial frequency line in its spectrum) is known as a
minimum redundancy array. Although many minimum redundancy arrays are also
multiplicative arrays, one simple form of such an array26 is shown in Fig. 13.27 and
corresponds to a square law detected additive array. This array consists of a
7-element linear array in which three elements are missing. The missing elements are
so chosen that the resulting combination contains all the possible inter-element
spacings of the full 7-element array. This means that its spatial frequency spectrum
must be uniform (with the exception of the DC self product term.) If this
4-element array is square law detected, the resultant directional pattern corresponds
to the sin 7p/sin p pattern of the full 7-element array plus the DC term due to the
four self product terms. This pattern is shown in Fig. 13.28 for both linear rectifi-
cation and square law detection.

Fig. 13.26 Synthesis of a 74-element array, 30dB sidelobe response using a 12 x 16 multipli-
cative array

13.4.5 Multiple source excitation of multiplicative arrays


It is common practice to assume that the output of an array to multiple sources
may be computed on the basis of linear superposition. This is, of course, true in
terms of the fields in space and the resultant signals prior to demodulation. How-
ever, the non-linear process of detection is not reversible and introduces cross
product terms between targets. For most filled linear arrays such cross product
366 Array signal processing
term are not very important unless it is necessary to examine details of target reso-
lution performance. In the case of minimum redundancy arrays or multiplicative
arrays the situation can be very different.27'28'29 Such terms can cause confusing
responses and their amplitude may be larger than sidelobe levels.

O O •+• • O • O

Fig. 13.27 ARSAC array: a 4-element minimum redundancy array containing all the inter -
element spacings of a 7-element array

1.00r

.- 0.75

2 0.50

0l—
-1.0 -0.5 0 0.5 1.0
sin 6

1.00r

0.75

£ 0.50

I
^0.25

-1.0 -0.5 0.5 1.0

Fig. 13.28 Directional patterns of 4-element ARSAC array.

If a directional receiver having a directional pattern D(6) is illuminated by a


source (or reflected signal from a target) of strength A x with frequency co and direc-
tion 0u the receiver output will be
Array signal processing 367

E = ^1Z)(i91)cos(ojzt + 01) (13.13)


where (pt is a general phase angle. If this filter signal is square-law-detected and low-
pass filtered, the output will be proportional to
E = \A\D\dd (13.14)
A similar output would occur for separate excitation of the receiving array with a
source A2 at angle d2. However, for simultaneous excitation of the directional
receiver, with both signals, the output after square-law detection and filtering
becomes
Ex = U2iD2(d1) + U22D2(d2) + A1A2D(e1)D(62)cos((t>1-(l>2) (13.15)
The first two terms are the outputs that would be predicted by superposition, and
the third terms is the cross product between two sources or targets.
Note that this cross product terms exists for all forms of directional receiver
(whether using continuous apertures, filled arrays, or thinned arrays) and for all
forms of demodulation. This cross product terms is an unwanted distortion, which
reduces the ability to resolve multiple targets; thus it needs to be reduced (ideally
to sidelobe levels) relative to the wanted target self-products. Cross products
between the signals from different element positions are the essential feature of a
directional receiver; it is the cross product between different targets that must be
removed.
For filled arrays with at least one of the sources in the sidelobes, the target cross-
product terms is reduced because of the low value of its product Z)(0i)Z>(02). This
fact shows that for filled arrays, such cross products would be reduced to sidelobe
levels if the angle between the two sources is greater than approximately one beam-
width. For such filled arrays the effect of the cross product terms is to modify the
two-target response of the array when both targets are within the main beam. This
leads to beam pointing errors. The amplitude of the target cross products is also
proportional to cos(0! — <p2), which represents the difference in phase of the
returns from the targets, A i and A2. The phase difference will depend upon the pre-
cise ranges of the 2 targets.
For radio astronomy, the signals from two different radio sources are uncorrela-
ted so that all such target cross-product terms average to zero, which explains the
extensive use of multiplicative arrays and thinned arrays in radio astronomy. How-
ever, for a radar system in which both targets may be illuminated from the same
transmission, it may be necessary to introduce some deliberate variation of
( ^ — 0 2 ) plus some averaging (integration) to integrate out the target cross
products when a low-redundancy arrays is used. Frequency agility of the radar
transmissions may be used to achieve this effect.

13.4.6 Signal and noise performance of multiplicative arrays


It will be seen from the above that the performance of any form of directional
receiver is determined not only by the angular distribution of the wanted signals
(and unwanted noise) but also by coherence properties of both signals and noise in
368 Array signal processing

angle and time. The multiple source cross products are reduced, and hence the reso-
lution properties of the receiver are enhanced by any lack of temporal coherence of
the wanted signals. In some situations this can occur naturally due to the ionos-
phere or multiple reflection fading.
A detailed discussion of the noise performance of multiplicative arrays and cor-
relation detection systems is outside the scope of this book. It has received atten-
tion from many authors.30'31 Correlation receivers are used extensively for passive
detection systems in several applications including radar and sonar.
If we consider the simple case of a uniformly phased multiplicative array it is
reasonable to ask whether the multiplicative signal processing affects the detection
process. This corresponds to comparing the multiplicative demodulation process
with other forms of demodulation. A simplified picture can be obtained from the
output SNR of a multiplier in terms of the SNR at its two inputs. Tucker has pro-
vided an expression for this in the form:30
2/? R
R ( 1 3 1 6 )

where R x and R2 are the SNRs of the two input signals. This may be compared with
a combined SNR of (Rt + R2) for optimum combination of a linear additive array.
A limitation of this approach is that the amplitude probability distribution of
the noise outputs are not identical to that of a simple detector.31 The probability
distribution is an exponential and Bessel function dependent upon the correlation
of the two inputs to the multiplier. Nevertheless calculations on probability of
detection performance reveal a loss of about 1-1.5 dB compared with conventional
detection. This small loss is therefore associated with the absence of the self
product terms at the output of the directional reciever. An important consequence
of the above equation is that the output SNR is a maximum when the two input
SNRs are equal. This means that the two sub arrays feeding the multiplier should
have equal gains.
In cases where thinned multiplicative arrays are used, the basic SNR at the two
inputs will be reduced by the thinning factor of the two sub arrays.

13.4.7 Post demodulation synthetic aperture antennas


It will be apparent from the former sections that, in order to determine the direc-
tion of a single radio signal, it is not necessary to employ a filled array and some
form of minimum redundancy array may be an attractive configuration. The radio
astronomy field has the notable advantage that all the multiple source cross-product
effects may be completely removed since the product of signals derived from two
different radio sources averages to zero.
We have shown that the output information from a line array may be broken
down into self product terms and cross product terms between different pairs of
antenna elements corresponding to inter element spacings of d, 2d, 3d,.... The
location of these cross product pairs need not even be located within the line of the
array, provided that the distance d (or nd) is measured in the direction of the array.
Array signal processing 369
32
This property has been discussed by Shearman, who drew attention to the
features of non-co-linear multiplicative arrays. Figure 13.29 shows two different
multiplicative arrays which are not co-linear but still have identical patterns (for
situations when cross products can be averaged out).

d, d2 d3
QJ-Q--Q Q Q O

Fig. 13.29 Two multiplicative arrays having identical directional responses despite the differ-
ent array geometry

For the case of radio astronomy, the far field distribution of sources is fixed so
it is not even necessary to take the output signals from all the elements at the same
period of time. This concept led to the development of synthetic aperture antennas
for radio astronomy by Ryle.2
The linear array synthetic aperture is shown in Fig. 13.30. It is necessary to have
two elements and one multiplier and to move the elements to different locations
along the line and to record the bipolar, low pass filtered output of the multiplier
for all the necessary interelement spacings corresponding to a filled array. These
recorded output signals can then be summed directly to form the signals corre-
sponding to a synthesised beam. If the output of the multipliers were also recorded
for in phase and quadrature phase of the two elements, it would also be possible to
synthesise multiple beams pointing in many different directions.
370 A rra y signal processing
Since the outputs from these multipliers are, in general, broad bandwidth signals,
they may be integrated in order to increase the signal to noise ratio. This provides
a narrow band signal which may be conveniently stored on magnetic or paper tape.
These recorded signals may then be fed into a computer to produce various forms
of synthesised beams at a later date.

positions for
fixed moveable element
element \

Y
I

Fig. 13.30 Synthetic aperture array (one-dimension synthesis)

This is the basis of the post demodulation form of aperture synthesis used in
radio astronomy. The term post demodulation is relevant because it is the demodu-
lated output of the multipliers which are stored and then used to synthesise patterns.
Although the above discussion has been in terms of a linear array, the technique is
quite general and could apply to a planar or even a volumetric array.
If et represents the output of the zth element of an ^-element array, the square
law detected output of the whole array Eo is given by:
2
En = (13.17)
£=1

This expression shows that the detected output can be synthesised by a series of
multiplicative products between the zth and gth elements. This also indicates that
this multiplicative system has an exact equivalence with the additive square law
detected array. This applies even to multiple target performance if the array is syn-
thesised to include all repeated element products.
To perform two dimensional aperture synthesis, the two elements must be loca-
ted at appropriate points in a plane, i.e. (x\9y.i) for element 1 and (x2,y2) for
element 2. The locations within this plane are then chosen such that x{ — x2 and
J i ~ y i cover all required interelement spacings in both dimensions. In many cases,
it is more convenient to employ more than two elements in order to reduce the
total amount of time required to synthesise patterns. Various combinations of
moveable elements and a larger fixed array may also be adopted.
Although the basic concept of synthetic aperture appears to lead to a very low
gain performance, the important parameter is the time taken to collect sufficient
energy to achieve an adequate signal to noise ratio for each direction in space. Here,
a synthetic aperture system scores on the grounds that it can also simultaneously
synthesise multiple beams in different directions, which therefore speeds up the
Array signal processing 371

overall process. The main advantage of the synthetic aperture antenna is its substan-
tial reduction in cost and its ability to synthesise antennas whose dimensions would
be quite impractical by means of a filled aperture technique.
An interesting application of post demodulation synthetic aperture techniques
to radar has been described by Shearman.33 This is an HF radar employing flood-
light transmissions covering a broad sector of interest and is designed to study the
ionosphere at long range.
To achieve adequate angular resolution at these frequencies requires a long array
and consequent expense, but the movement of the targets is quite slow so that
there is ample time to build up a picture. The receiving array consists of a 2-element
multiplicative interferometer having one fixed element and the other moveable (on
rails). The product of these two element outputs is therefore recorded for-various
positions of the moveable element and group of narrow beams are synthesised
covering the sector illuminated by the floodlight transmitter. Again, since the
recorded signals are not dependent upon the phase of the received signals, there is
no need to employ coherent transmissions. An important restriction on the tech-
nique is that if the target field is complex (as it is for ionospheric targets), there will
be a significant distortion of the results due to multiple target cross product effects.
A particularly interesting result of measurements made by Shearman on his radar is
that the effects of such cross products could be substantially reduced by the appli-
cation of frequency agility to the transmissions.34 This indicates that the use of
frequency agility should be able to reduce such cross product effects when mini-
mum redundancy arrays are employed in a radar system.

13.5 Coherent synthetic aperture antennas

13.5.1 Principle
The technique of coherently synthesising an antenna aperture is primarily associa-
ted with airborne or satellite-borne mapping radars, as illustrated in Fig. 13.31. The
basic principle is shown in Fig. 13.32. A pulsed radar transmitter feeds a compara-
tively small antenna aperture d, which is transported along a straight line. Received
signals are range-gated and, for each range interval, the amplitude and phase history
is stored. The processor associated with each range interval performs a continuous
weighted summation of signals over a distance D equal to the length of aperture
which it is desired to synthesise. Thus the real aperture d essentially forms the ele-
ment pattern for the synthesised aperture D.
The system differs from the post-demodulation synthetic aperture antennas dis-
cussed previously in three major respects:
(a) The transmitter provides the phase reference, so that absolute phase informa-
tion is stored, rather than relative phase between a pair of elements. The coherent
synthetic aperture antenna is thus only suitable for detecting stationary, non-
scintillating targets.
372 Array signal processing

V = 100 m/s

V = 7500 m/s

, =880 km
Ro =840 km
h = 800 km

Fig. 13.31 Principal application areas for synthetic aperture radars


a Typical airborne mapping role
£ Typical satellite-borne mapping role
Array signal processing 373

(b) Since 2-way phase is involved, angular discrimination is generally twice as good
as that of a post-demodulation synthesised aperture.
(c) Targets of interest are invariably within the near-field range of the maximum
aperture which can be synthesised.

processed output
E(x)

Fig. 13.32 Principle of coherent synthesised aperture antenna

Assuming that the radiation pattern of the real aperture d can be approximated by
a flat-topped beam of angular width \/d, then it is evident from the geometry of
Fig. 13.32 that the maximum length of aperture which can be synthesised as pro-
portional to range, specifically:
(13.18)
374 Array signal processing
For the airborne radar example of Fig. 13.31a, typically d = Im and X = 0.03 m,
and Do becomes 200m at Ro= 10km. The far-field range of such a synthesised
aperture, according to the usual D%fk criterion, is 3000 km. For the satellite
example of Fig. 13.315, typically d = 10m and X = 0.25 m, giving Do = 22km at
880 km range. The far-field range in this case is about 2 x 106km.
When the aperture d has moved a distance x along the track, the additional path
to the target is
8 = [R2 + x2]QS-R » x2/2R (13.19)
and hence the phase variation is
0 = - 4 T T S / X ** -2TTX2/RX (13.20)

The received signal can therefore be represented by


f(x) = A(x) exp [-j2nx2/R\] (13.21)
where Aix) includes the effects of the beam shape of the real aperture d and any
target scintillation effects. For a flat-topped beam and a point target:
fix) = exp [-J2<rrx2/R\], (-D0/2<x <D0/2) (13.22)
The parabolic nature of the phase history is shown in Fig. 13.33. The maximum
phase excursion is directly proportional to range and inversely proportional to the
square of the real aperture d. The distance along track for which the phase devia-
tion is less than 7r/2 is
1 = y/(RX) (13.23)
which can be recognised as the fraction of the synthesised aperture having a far-
field range just equal to R.
The processor performs a weighted summation of signals received over a distance
D which generally lies between L and Do depending, inter alia, upon the amount of
processing power available. Assuming for the moment that the sampling is virtually
continuous over this distance, then the processed output can be approximated by

E(x) = W(y)f(x~y)dy (13.24)

where W(y) represents the amplitude and phase weighting applied in the processor.
Essentially, the processor associated with each range interval performs a correlation
between the range-gated received signals f(x) and the stored weighting function
W(y).
The principal parameters under the designer's control are the real aperture length
d, the synthesised aperture length/) and the processor weighting function W(y).

13.5.2 Unfocused synthetic aperture


Early synthetic aperture radars operated in the unfocused mode, in which the proces-
sor provides constant phase weights so as to simulate a constant phase illumination
Array signal processing 375

over the synthesised aperture. In this case, it is evident from Fig. 13.33 that it is not
worth while choosing D to be much longer than L, since the signal contributions
outside this interval are largely self-cancelling due to the rapid phase variations.

phase

TTRA

=L
distance along track
(a)

phase

=L
distance along track •
Fig. 13.33 Phase history of target (b)
a Target at short range
b Target at longer range

For the simplest case of constant amplitude and phase weighting, then

(13.25)
and the processor output, using eqns. 13.22 and 13.24, becomes:

E(x) = JQ exp[~~j2n(x-y)2/R\]dy (13.26)

which reduces to
376 Array signal processing

E(x) =
2
(fovD-D0l2<x<D0/2-D) (13.27)
where
F(u) = C(u)-jS(u) (13.28)

and C(u) and S(u) and standard Fresnel integrals defined by:

and S(u) =Jo" sin {^12) dv (13.29)

The form of the processed output is illustrated in Fig. 13.34.


The peak response occurs at x = D/2. For D less than X, the net phase variation
in eqn. 13.26 is substantially linear and the output waveform is of (sinz)/z shape
with a — 3 dB angular width of X/2Z), giving an along-track resolution of A/?/2D =
L2/2D. The peak signal amplitude is proportional to D. For D equal to /,, the effect
of the parabolic phase variation becomes noticeable, and the minima near the main
response become filled-in. For D greater than L, the peak signal amplitude does not
increase and the waveform exhibits the typical wide beam response of a de-focused
aperture, with substantial Fresnel-type ripples.
It will be appreciated that the noise power from the processor is directly pro-
portional to D. The condition D = L thus represents an optimum choice of synthe-
sised aperture length for both peak signal-noise ratio and for resolution. The along-
track resolution of an optimised unfocused synthetic aperture antenna is therfore
L/2 = y/(RX/4) (metres) (13.30)
and the corresponding angular beamwidth is
B = \J2L = y/QQW) (radians) (13.31)
This is half the normal beamwidth of an aperture of length L, due to the 2-way
phase effects.
In a practical system, the weighting function W(y) includes appropriate ampli-
tude weights which, together with any amplitude weighting due to the beam shape
of the real aperture d, can be chosen to improve the sidelobe response. The usual
trade-offs between peak signal level (cf. antenna gain), sidelobes and beamwidth
apply.

13.5.3 Focused synthetic aperture


A dramatic increase in resolution is obtained by operating in the focused mode,
where the weighting function includes phase terms which compensate for the
Array signal processing 377
parabolic variation in signal phase. Thus for each range gate the weighting function

W(y) = exp[jir(D-2y)2/2R\] (13.32)


plus any amplitude weighting which is introduced to control the sidelobes

a/— D/2 U-
"U
B i 1
mli D=T/A R
D

/AR 3/AR

A
\
%
B
/ \ D-
pli

/XR

/ V.
E
o

-/AR /TR 2/"AR 3x/TR

D=2/XR

-/AR

distance along track -^


Fig. 13.34 Processed output from unfocused synthetic aperture

An indication of the resolving power can be obtained by considering the size of


the focal region for which the phase error is less than TT/2. In Fig. 1335a it is
assumed that the processor is set up to compensate the path lengths for the target
at 70, e.g. the paths Rt and R are equalised. The net path error between rays from
the centre and from the edge of the synthetic aperture D to the target at Tx is:
5 = {zl-Z)-{R1-R) (13.33)
or 2 2
8 * (nD/2R) + (iD !8R ) (13.34)
378 A rray signal processing

track

(D ft /D)* (

track

(b)
Fig. 13.35 Focal region in a focused synthetic aperture antenna
a Geometry
b Focal region
Array signal processing 379

where 77, £ are the along-track and across-track co-ordinates of Tx from To. Setting
6 = A/8, corresponding to n/2 2-way phase error, gives the maximum values of dis-
placement as
7? = ±(XR/4D) = ±(dl4)(D0/D) (13.35)
and 2 2 2 2
% = ±(XR /D ) = ±(d IX)(DolD) (13.36)
The focal region is roughly elliptical, as indicated in Fig. 13.35Z?. For an optimised
focused synthetic aperture D =D0 and the along-track resolution is simply d/2, i.e.
half the aperture of the real antenna, whilst the across-track resolution is 2d2/X,
which is just twice the Rayleigh range of the real aperture. In practice, the across-
track resolution is enhanced by employing short pulses and range-gating.
The apparent anomaly that the along-track resolution is directly proportional to
d, rather than inversely proportional to d as in a conventional antenna, is resolved
when it is recognised that choosing d small enables a larger value of synthetic aper-
ture d to be realised. The penalty paid for increasing the resolution is that more
processing power is required.
The waveform of the processed output may be obtained by substituting eqns.
13.32 and 13.22 into eqn. 13.24. For the case of uniform amplitude weighting and
an optimised system (i.e. D = Do), the processed output becomes:

E(x) = \ exp[jn(D0-2y)2/2RX]Qxp [—/2TT(X -y)2/R\]dy (13.37)


or
E(x) = exp \Jir(Pl- 4x2)/2RX] J exp [-j2ny(po-2x)lR\]dy (13.38)

where the appropriate integration limits, from eqn. 13.22, are


y = 0to(D 0 /2+jc) for-D o /2 < x < A > / 2
corresponding to the 'charging' interval of the processor, and
y = (x - Do/2) to Do for D o /2 < x < 3D0/2
corresponding to the 'discharging' interval.
The phase term inside the integral of eqn. 13.28 represents a linear variation
with y, the distance along the processor, with a slope proportional to the along-
track distance x. Upon performing the integration, the result is:

This can be expressed more compactly by shifting the origin to the position of peak
response by writing:
X = x-D0/2 (13.40)
380 Array signal processing

Then, noting that DQ = (RX/d), the final expression for the output from the pro-
cessor becomes:
(D<X<D) (13.41)

-Do/2 3Do/2
distance along track-•— x

Do/2)
distance along track —«*•• x
b
Fig. 13.36 Processed output from focused synthetic aperture
a Optimised {D = Do)
b Non-optimised (D < DQ)

As illustrated in Fig. 13.36, the peak output occurs at X = 0 (x =D 0 /2), with an


amplitude of DQ and a — 3dB width of dj2. The waveform is substantially of
(sinz)/z shape, but the periods of the sidelobes increase as the distance from the
central peak increases, due to the reduction in the length of signals stored in the
processor. The total duration of output signals is 2DQ.
For a non-optimised focused synthetic aperture (D<D0), the peak signal is
reduced to D and the beamwidth increases to (dD0/2D). The waveform is still sub-
stantially of (sinz)/z shape and the total signal duration is (Do+D).
Sidelobes can be improved in the usual way by introducing amplitude taper into
the processor weighting function W(y).

13.5A Minimum number of samples


The formulas derived above are based on the assumption of continuous sampling
along the synthetic aperture. In order to accurately reproduce the phase histories
Array signal processing 381

Illustrated in Fig. 13.33 from a finite number of samples, the maximum phase shift
between samples should not exceed n. The maximum phase slope occurs at the
edges of the beam from the real aperture, and at these points the along-track sampl-
ing interval s corresponding to a phase shift of n is d/2. In practice, a shorter
sampling interval may be needed, to allow for the fact that the beam from the real
aperture d extends beyond the assumed beamwidth of \/d.
Even when the synthesised aperture D is less than the maximum possible value
of D o , it is still necessary to sample at intervals of d/2 to prevent aliasing effects
(e.g. a short section of the parabolic phase curve at the edges of the illuminated sec-
tor could appear to have a reduced phase slope if it is under-sampled).
It will be apparent that the sampling interval s determines the angular positions
of grating lobes from the synthetic aperture. As shown in Fig. 13.37, grating lobes
occur at intervals of approximately X/2s, and it is clearly desirable that the beam
from the real aperture d should have low side lobes at these angles.

sin 8 = A /2s

h-.-l
sampling
interval
Fig. 13.37 Grating lobes in a coherent synthesised aperture

The minimum number of samples for the processor for each range interval is
M = 2D/d = (2R\/d2)(D/D0) (13.42)
The number of range gates is
N = {Rx~R0)jbR (13.43)
where bR is the range gate width and Ro and Rx are the minimum and maximum
instrumented ranges. The minimum total number of samples or pixels in the proces-
sors of an optimised focused synthetic aperture radar is therefore
382 Array signal processing
P = (K/d^iRl-RDldR (13.44)
For the airborne radar example of Fig. 13.31a, where, say,/?! = 10km,i? 0 = lkm,
d = lm, dR = lm and X = 0.03 m, then P is 3 x 106. For the satellite example of
Fig. 13.316, typically Rx = 880km, Ro = 840km, d= 10m, 8R = 10m and
X = 0.25 m. The number of pixels in the processor is then about 2 x 107.
Although the along-track resolution is independent of wavelength, choosing a
short wavelength helps minimise the amount of processing power required.

13.5.5 Signal-processing hardware


Early airborne systems simply recorded the raw data on photographic film. The
signal processing was ground-based, generally employing coherent optical tech-
niques35 to perform the correlation-type processing defined by eqn. 13.24. More
recently, fast compact on-line digital processors have become available,36 making
it possible to produce maps in real time whilst airborne.
Pulse-compression techniques are often used for the radar to achieve fine range
resolution. The processor then usually incorporates the algorithm for recovering
the compressed pulse in addition to performing the along-track processing.

13.5.6 Radar PRF


The PRF of the radar is related to the sampling interval s and the vehicle velocity V
by
p = V/s (13.45)
Choosing a sampling interval ofd/2 gives:
Minimum PRF = 2 V/d (13.46)
The maximum PRF is related to the maximum unabiguous range bracket (Ri~RQ)
by:
Maximum PRF = c/2(R1-R0) (13.47)
where c is the velocity of the electromagnetic waves.
For the airborne radar example, with d = lm, V= lOOm/s and (R1—Ro) =
10 km, the PRF limits are 200 Hz to 15 kHz.
In satellite applications, the choice of PRF is usually much more restricted, since
the vehicle velocity V is much greater. Thus d = 10 m and V = 7500m/s results in
a minimum PRF of 1.5 kHz. The maximum PRF for the example of Fig. 13.316,
determined by the nominal unambiguous range bracket of 40 km, is 3.75 kHz. In
practice, this maximum has to be reduced substantially, since the elevation beam
will intercept targets at ranges below and beyond the nominal bracket.
13.5.7 Doppler-frequency interpretation
System designers usually refer to the parabolic phase variation given by eqn. 13.20
as a linear change in Doppler frequency as the real aperture sweeps past the target.
The frequency, obtained by differentiating eqn. 13.20, is
/ = -(l/2ir)d<p/dt = (2x/RX)dx/dt = 2Vx/RX (13.48)
Array signal processing 383

Setting x = ±DQj2 gives the maximum Dopper frequency as:


Maximum/ = ± V/d (13.49)
The spread in Doppler frequencies is thus equal to the minimum PRF given by eqn.
13.46. This value represents the nominal — 3dB spread. The total spread is some-
what greater, since the beam from the real aperture extends beyond X/d. If the
minimum PRF is chosen, these wider-angle Doppler spectrum components are
folded over into the same band as the main signals.
The signal processing to achieve the along-track range resolution is analogous to
the recovery of 'chirp' signals from a radar employing a swept-frequency transmit-
ted pulse. The bandwidth in the present instance is of the order of a few hundred
hertz, compared with tens of megahertz for chirp, whilst the total signal duration is
perhaps a few seconds, compared with microseconds.
The grating lobe angles shown in Fig. 13.37 can be interpreted as angles of am-
biguous Doppler, since the angular spacing is
X/2s = pl(2V/\) (13.50)
i.e. the angular spacing is equal to the PRF divided by the Doppler frequency cor-
responding to the vehicle's velocity.

13.5.8 Radar performance


The single-pulse signal-to-noise ratio may be obtained from the usual radar equa-
tions.37 The received power per pulse is

Pr = ^ ^ (13.51)
' (47T)3i?4
where Pt is the peak transmitter power, G is the gain of the real antenna aperture d,
and a is the target echoing area. The noise power is
Pn = kTBF (13.52)
where k is Boltzmann's constant, T the receiver temperature, B the bandwidth and
F the noise factor. B is generally approximately 1/r where r is the pulse length.
The vertical beamwidth of the antenna is fixed by the coverage requirements,
and hence the gain G is proportional to d/X. The single-pulse signal-to-noise ratio is
thus independent of the choice of wavelength, but is proportional to d2.
The quality of the processed image may be judged by its resolution, its ability
to reject noise and its ability to distinguish wanted targets from the background sig-
nals contributed by other targets. The resolution, of course, is determined by d and
r.
All samples are integrated coherently; hence:
S/Nimprovement factor = D/s = (RX/sd)(D/D0) (13.53)
less any loss due to amplitude taper introduced to improve the sidelobes (which is
directly analogous to antenna efficiency). For an optimised focused synthetic
384 Array signal processing
aperture, sampled at the minimum PRF, eqn. 13.53 becomes:
S/Nimprovement factor = 2D0/d = 2RX/d2 (13.54)
which is 28 dB and 36 dB at the maximum ranges of the airborne and satellite-borne
examples quoted previously. The effective S/N at the output of the processor
becomes independent of d, but is directly proportional to wavelength and inversely
proportional to R$. Reducing the sampling interval s by increasing the PRF above
the minimum value will improve the processed S/N, but only at the expense of
more processing power.
The contrast between a wanted target and the residual signals from other targets
within the processor depends upon the number of targets being processed simultan-
eously and the sidelobe level achieved in the processor. If S represents the average-
power sidelobe level relative to the peak response, then in the case of an optimised
focused system where the wanted signal duration is d/2 and each resolution cell
produces residual signals over an interval of 2D0-
Contrast factor = d/(4SD0) (13.55)
Thus the contrast factor is essentially half the peak signal-to-mean-sidelobe ratio,
degraded by the S/N improvement factor. The need for low sidelobes from the pro-
cessor is quite apparent. For an under-optimised focused system:
Contrast factor = (dD0)l[2SD(D0+D)] (13.56)
Reducing the sampling interval below d/2 will not affect the contrast factor, since
both wanted and unwanted signals are correlated to the same degree within this
interval.

13.5.9 Response to target scintillation and motion


Any target movement will destroy the phase coherence and in general results in a
reduction in the S/N improvement factor, the surplus energy being smeared over a
number of resolution cells.
Even with completely stationary targets such as buildings, there may be several
scattering centres within the resolution cell, and this can give rise to the signal
amplitude and phase being aspect dependent, resulting in some smearing. In order
to provide better images in these circumstances, the length of aperture which is
coherently synthesised is sometimes reduced, but a number of such apertures
ahead of and behind the normal beam are synthesised by including appropri-
ate linear phase shifts in their processors. The video outputs from the various 'sub-
apertures' are then added together non-coherently.
For terrain targets with waving vegetation, the speeds involved are generally a
few metres per second (corresponding to Doppler frequencies of around 100 Hz at
X = 0.03 m). If the Doppler beam width of the processor is of the same order, as is
often the case for airborne systems, coherence is destroyed and the processed out-
put from adjacent pixels takes on a noise-like appearance.
A moving point target may traverse several across-track range cells and produces
Array signal processing 385

an additional linear variation of phase with time (i.e. a fixed Dopper offset) within
each cell. This generally results in along-track translation errors, but the target may
be blind if the phase shift between samples is exactly rt. For moving vehicles, there
is further spectrum broadening due to aspect dependency, vibration and multipath
effects.

13.5.10 Applications
The two principal application areas are sketched in Fig. 13.31. It will be evident
that deviations of the flight path from the desired straight line must be assessed
and compensating phase adjustments made to the processors. The compensation is
range-dependent, because the depression angle to the target varies with slant range.
Additionally, the real antenna's beam must be stabilised against vehicle pitch and
yaw, unless its beamwidth is very broad. The residual errors caused by these effects
can be separated into random and periodic errors and treated by the usual methods
for assessing the effects of antenna illumination errors.38 Additional phase errors
will arise if there are any spatial or temporal variations in the refractive index of the
propagating medium.39
Airborne mapping, as in Fig. 13.31a, was initially developed for military surveill-
ance, but is finding increasing scientific applications, particularly in remote sensing
of the terrestial environment.40
A number of satellite-borne synthetic aperture radars are currently operating for
both ground and oceanographic surveillance.41 They usually orbit at a compara-
tively low altitide (c. 800 km) and their processors require additional compensation
for the effects of earth rotation.
Since the basic resolution is independent of wavelength, synthetic aperture
radars can be built at VHF. Such low-frequency systems are valuable in remote
sounding applications in glacial ice42 and dry soil,43 as deep penetration can be
achieved due to the low absorption coefficients at these frequencies.

13.6 Null steering

13.6.1 Introduction
Communications and radar antenna systems are susceptible to degradation in per-
formance because of interference received through their sidelobes. The interference
may be deliberate electronic counter measures (ECM), accidental RF interference
(RFI), clutter returns or natural noise sources.
Typical interference signals received by RF systems are from point sources
radiating a coherent wavefront for at least the signal gathering time of the antenna.
The antenna operating in its 'far field' will receive a plane coherent wavefront. In
order to prevent interference of this type entering the receiver it is necessary to
arrange for a null in the antenna's polar response to coincide with its angle of
incidence.
The existence of a plane coherent wavefront can be identified through the use of
386 Array signal processing
an array of receiving elements. Once identified, the angle of incidence of the signal
can be readily determined. The directional pattern of an iV-element linear array may
possess up to (JV— 1) independent zeros. There are obvious advantages in having
the ability to independently steer the zeros in the array's directional pattern in
order to reduce interference signals normally received by the sidelobes.
In order to independently control the angular location of the zeros it is
necessary to control both the phase and amplitude of signals in the feed system of
each element. The requirement for accurate control of both the phase and ampli-
tude of signals received by each element can in practice be difficult to achieve, and
often limits the use of such systems to experimental and military applications.
Before discussing the properties of zeros further it is worth defining the differ-
ence between minima (commonly called nulls) and zeros. The mathematical analysis
of an ideal array will define angles at which the response is zero. All incoming
signals at those particular angles of incidence and at a single defined frequency will
not produce any output. However, because the formation of zeros relies on many
signals cancelling exactly in the combining network, they are extremely sensitive
to aberrations in the array, and to the bandwidth of the signals. Even if it were
possible to manufacture a perfect antenna, it is likely that mutual coupling effects
would disturb the signal phase and amplitude characteristics. Therefore in practice
zeros are not perfectly formed. They will not have infinite attenuation at one
angle, but a rather less sharp response spread over a range of angles. Mathematical
zeros therefore tend to produce nulls in the actual polar response of an antenna.

13.6.2 Multiple Nulls


Schelkunoff^ showed that the directional pattern of an iV-element linear array may
be expressed as a polynomial of order (N— 1). Hence for the array shown in Fig.
13.38 the polynomial is
Wx + W2z + Wsz2 + WNzN~l (13.57)

=I k=
where *
z = exp (j(2ird/X) sin 6) (d = element spacing)
and the phase reference is taken as the first element in the array. The coefficients of
the polynomial may be complex, thus representing the amplitude and phase weight-
ing of each array element. Thus:
Wk = Akexp(j<Pk) (13.58)
The polynomial may be factorised into (N — 1) roots and each of these represent a
zero of the polynomial
D(d)= WN(z-b1)(z-b2)(z-b3)...(z-bN-1) (13.59)
In general, these roots are complex numbers and may therefore be represented as
points in the complex z plane. Since z is simply a phasor, \z\ = 1; so roots located
Array signal processing 387

on the unit circle \z\ = 1 represent true zeros of the directional pattern (i.e.

This Schelkunoff analysis also shows44 that if d > A/2 the zeros of the pattern
function may repeat over the 180° of real angles and if d < A/2 there may be zeros
of the pattern outside the range of real angles. Although the above indicates the
maximum number of independent zeros to be (N— 1) there is no reason why a pat-
tern may have no true zeros if all the roots are located off the unit circle. Mutual
coupling and tolerances of excitation can give errors in the coefficients of the poly-
nomial for a real array. This displaces the zeros off the unit circle and hence softens
the otherwise sharp nulls in the pattern.

w2 W
3

Fig. 13.38

The relationship between the zeros, the radiation pattern and the element
weights W is sometimes difficult to visualise. Consider the pattern of the uniformly
illuminated array of Fig. 13.39 and assume that there is a jammer in the sidelobe
shown. It would be sensible to move the zero closest to the jammer on to the
jammer bearing. In z-plane terms (eqn. 13.59) this is easy. The root bp correspond-
ing to the closest zero is simply replaced by Bj, where
. 2nd
exp (/-—sin0;- (13.60)
A
and 0j is the jammer bearing. The modified radiation pattern, shown in the Figure,
can be expressed as

Dj(6) = (13.61)
388 Array signal processing
and it is clear that the disturbance of the pattern is mainly in the region of the jam-
mer. However, the effect upon the element weights W is quite complicated and it
will be appreciated that every weight must be modified in order to achieve the new
pattern. The situation becomes more complicated if there is more than one jammer
to be dealt with.

•i
o
<L

-30
-180

Fig. 13.39 Effect of moving one zero on the pattern of an 8-element uniformly illuminated
array

13.6.3 Multi-element null steering arrays


A network capable of independently steering all (TV — 1) real zeros of an TV-element
linear array has been described by Da vies.45 This is shown in Fig. 13.40. The first
row of identical (ganged) phase shifters combine elements in adjacent pairs and
hence results in steering a single zero. The pattern of each pair of elements will be:

= cos (13.62)

There are three outputs with this directional pattern. These are then combined
together again in adjacent pairs with a phase shifter 0 2 to steer the second zero, and
the process repeated for the third. For an TV-element array, the overall pattern is the
product of (TV' — 1) such terms:

D(d) cos {( — %md\ — j 0 r (13.63)


r=l

and it will be appreciated that all the zeros are real zeros on the unit circle.
Fig. 13.41 shows a sidelobe region of a directional pattern with zeros at Un and
£/(n+1). If the two zeros are made coincident a rather broader second-order zero
Array signal processing 389
results where the pattern touches the axis. This can be extended to third order and
higher order zeros as shown in Fig. 13.42.
Multiple zeros deepen and broaden the null, and as a consequence reduce the
level of the lobe between nulls. In cases where interfering signals are received by the

array

output
F ig. 13.40 Null Steering Tree
H = Hybrid (4th port terminated)

array from specific directions, multiple zeros are useful in preventing the unwanted
signals entering the receiver. However, it is not always advantageous to overlay nulls
on existing nulls. Superior performance can often be achieved by locating zeros
between existing ones, especially when signals with a wide bandwidth are being
received.
Although the above concept can, in principle, synthesise any directional pattern
by the choice of the location of the zeros (which in turn defines the roots of the
390 Array signal processing

polynomial) the system is not loss free. This is due to the power lost in the resist-
ive loads. The system also has the disadvantage that movement of zeros can also
result in changes in the gain of the resultant pattern in a 'specified main beam'
tion. This may be overcome by extending the concept discussed in reference 46
which uses the weighted combination of a sine and cosine directional response to
steer a zero without changing gain on axis. A modified form of null-steering tree
based upon this concept is shown in Fig. 13.43 for a 3-element array.

D(6)

Fig. 13.41

1st order
D(6)
2nd order

•3rd order

Fig. 13.42

For the arrays shown in Fig. 13.43 the nulls are steered by the variable gain
amplifiers giving a null for row m in direction dm due to the amplifier gain Rm:

Rm = cot p»sind (13.64)

The proximity of a null to the broadside is determined by the maximum allowable


Array signal processing 391
value of Rm. If Rm is in the range 1 to 0, i.e. a variable attenuator, and d = TT/2,
nulls can be steered in the range n/6 to TT/2 radians from the broadside. There is no
restriction on the angular spacing of the nulls, and closely spaced nulls do not repre-
sent superdirective directional patterns unless the interelement spacing is less than
X/2.

output

Fig. 13.43 Two-zero null steering array

13.6.4 An orthonormal beam-forming network


A rather different type of multiple null steering system has been discussed by
Nolen47 and by White48'49 and is based upon forming a complete set of orthogonal
beams. This is shown in Fig. 13.44. Each control element comprises a phase shifter
and variable power coupler, perhaps of the type shown in Fig. 13.45. If these ele-
ments are assumed lossless, then all the input power incident on terminals A and B
is shared between terminals C and D. This is also true for the whole array; i.e. all
the power incident at the elements is shared amongst the beam ports. The network
is therefore orthonormal. Whatever settings are made to the controls, the resulting
beams are always orthogonal and form a complete orthogonal set.
392 Array signal processing

couplers phase shifters

power patterns
H-i „ J A
spatial
filter
formed by
beam 1
spatial filter
formed by
beam land 2

spatial filter
formed by
beams 1,2 and 3 K/VJ _
F ig. 13.44 Orthogonal beam network

I Bexp(jv b )

phase shifter

B exp j (y*0)
A exp(jy a ) | I

I
3dB hybrid

variable coupler
ganged j 7 2*c (power coupling = sin ex)
phase shifters

3dB hybrid

!
i i C = cos o((A exp v a )
I +sinaf(Bexp \y+<t)

D=coscx(B expj vb*0) - sin a(A exp v a )

Fig. 13.45 Schematic of variable coupler used in orthogonal beam network


To align phases: 0 = {\jja — \jj &)
For zero output at D (after phase alignment): tan a = IBI/IAI
Array signal processing 393
A method of setting the controls to achieve 'open-loop' null steering in a jam-
ming environment is illustrated in Fig. 13.44. Here the phase shifters in row 1 (i.e.
0i2> 0i3 • •) and the values of the coupling coefficeints are chosen such that a maxi-
mum gain beam is produced in the direction of the first jammer, Bx. This beam
abstracts all this jammer's power and acts as a spatial filter for the next beam. Pro-
vided the first row of controls is undisturbed, any setting of the second row of con-
trols provides a beam with a null at 0v This row can therefore be set to provide a
maximum realisable gain beam in the direction of a second jammer. Note that the
full gain of the array will not be obtained unless 62 happens to be at a zero of the
first beam. Furthermore, the settings for the second row of controls are complica-
ted by the fact that they have to allow for the settings of the first row. This process
can then be repeated for the succeeding rows.
In an A^-element orthogonal beam network the first / rows of controls could be
set to abstract all the power from / jammers. The remaining TV — 1 —/controls
could then be set to provide multiple beams for desired signals. The interdepend-
dence of the control elements reflects the orthogonal beam properties. There are
essentially N— 1 degrees of freedom (or zeros) available for the first beam. Once
this has been set up, there are N— 2 degrees of freedom for the second beam and
so on.
Examples of the control settings are shown in Figs 13.46 and 13.47 for an
identical environment in which it is assumed that there are jammers at —90°, 45°
and 120° in sin 6 space. The only difference between these diagrams is the order in
which the beams have been set up. In both cases, the first beam set up is of classical
(siniVx/sinx) shape, but the other beams are affected by the spatial filtering effect
of the first beams. It will also be observed that beam 4 is identical in both cases: the
overall spatial filtering achieved did not depend upon the order in which the beams
were set up. This is because each beam is simply controlling one zero in the pattern
of all succeeding beams.
Calculation of the control settings required is straightforward, but tedious. If the
first beam is to be uniformly illuminated, then the power couplers in the first row
are set at 1/2, 1/3, 1/4 etc, and the phase shifters at 0, 20,30 etc. For the second
row, however, it is necessary to calculate the effect of the first row of controls
upon the signals arriving at the second row (for the second chosen angle of arrival)
in order to obtain the correct settings. This can always be done by working through
progressively from the left-hand side. The same procedure can be followed if
tapered illuminations are required.
Although the network as described would hardly be practicable for open-loop
control in a large array, its orthonormal properties are of fundamental importance
in the theory and, in some cases, the practice of adaptive antennas. This topic is
taken up later when eigenbeam analysis is discussed.

13.6.5 Multi-element array with single null steering


It is not always necessary to steer all of the (N— 1) nulls. When all elements are
identical and no amplitude tapering is used it is possible to simplify the array in Fig.
394 Array signal processing

13.38. A scheme which will allow a single zero to be steered in an array of any
number of elements is shown in Fig. 13.48. The single zero is controlled by combin-
ing signals from the two end elements. As in the multiple null steering array, the

beami
(-90°)

beam 2
(45*)

beam 3
020°)

beam 4

r\
11! \; w \..\ .
V

-180

Fig. 13.46 Network constants and radiation patterns for maximum gain beams at —90°, 45°
and 120°

control-amplifier gain is given by eqn. 13.64, the spacing d being the interelement
spacing. The position of the zero is subject to the same restrictions as in the mul-
tiple zero steering array, and its proximity to the broadside position is set by the
maximum allowable value for R. As a zero is steered close to the broadside, the
Array signal processing 395
position of the main beam changes. The main array response and the weighted end
element response of a 10-element array (constructed as shown in Fig. 13.48) is
shown in Fig. 13.49. This gives a zero at — TT/6 radians from the broadside position.

0,5 }—j 0.333 f ~ f 0.25

1075*

beam 2
(45°)

beam 3
-90°)

-180* -90°
2nd sin 0

Fig. 13.47 Network constant and radiation patterns for maximum gain beams at 120°, 4$°
and -90°

13.6.6 Sidelobe canceller


The technique for steering a single null discussed in the previous section is a special
case for identical elements and no amplitude taper. A more general approach is the
simple sidelobe canceller. This is shown in the schematic diagram of Fig. 13.50.
Here the output of a directional antenna is combined with the signals from a single
396 Array signal processing

omnidirectional element. If a zero is required in direction 0X it is necessary for the


signals from the omnielement to have the same strength as those from the direc-
tional antenna (in direction 0x) and to be in phase opposition.

V V V

Fig. 13.48 Multi-element single-null steering array

It is therefore necessary to vary both A x and 0j in Fig. 13.50 to steer this null. In
order to steer such a zero in the sidelobe region, it is necessary to subtract the
weighted omnisignal from the output of the array. This can also lead to a small
reduction of gain of the main antenna. In order to steer a null in the main lobe
region it is clearly necessary to attenuate the signal from the main antenna (or to
amplify the signal from the omni) prior to combination. This leads to a significant
loss of gain (or SNR of the received signal).
An alternative way of achieving sidelobe null steering in radar systems is to
detect the signals from the two antennas and then to subtract the demodulated sig-
nals.50 This is a form of post demodulation processing and it is necessary to take
care to avoid multiple signal (or multiple target) cross product effects. A simple
example of the above concept is merely to ensure that the detected response of the
omni element is larger than the largest sidelobe of the directional antenna. The sub-
tracted pattern thus has sidelobes, all of negative polarity. This has been termed
'sidelobe blanking'.
With this system a single signal received from the sidelobe region gives a negative
polarity output - which can be ignored (or removed with a diode). The perform-
ance in a multiple target situation, however, is unattractive since the negative signal
can still mask a wanted main beam return (which gives a positive DC polarity out-
put). Such systems are used in radar for the removal of impulsive interference.
Array signal processing 397

Sidelobe cancellation is therefore usually performed prior to demodulation.


There are many variants on such schemes51 and it is also possible to extend them to
multi-loop cancellers. These can steer N zeros by the use of TV separate auxiliary ele-
ments. This is discussed later in relation to the concept of adaptive sidelobe cancel-
lation.

polar component of 2 E-N+J/2 (E^+EQ

null steered t o - 3 0 ° polar component of j/2(Ej-Ein)


i.e. cot (ol/2) =-1

Fig. 13.49 Polar components of a 10-element null steering array

13.6.7 Discussion
The previous section has indicated how the zeros of various arrays may be steered
by combinations of amplitude and phase control. Although the discussion has cen-
tred on linear arrays, it may clearly be extended to planar or even volume arrays.
The general result of (N— 1) zeros applies to TV elements in any 3-dimensional
configuration. The particular case of circular arrays is dealt with in Chapter 12.
Although many of the networks discussed (such as the Davies tree) have indi-
vidual control parameters for each zero, the array excitation can, in principle, be
398 Array signal processing

expressed by 2(7V— 1) variables. These comprise the amplitude and phase of


(N— 1) elements compared to one element taken as reference. In many computer
control systems for null steering, it is therefore often more convenient to control
the amplitude and phase of each element directly and use the computer to work
out the complex changes of excitation needed for each movement of a null in the
pattern. This type of array will be analysed in the next few sections.

directional
antenna
_omnidirectional
"element

amplitude and
phase control

Fig. 13.50

In many practical arrays the presence of mutual coupling causes departures from
the desired directional response. Although this can be significant for many beam
patterns, it is critical for null patterns where the null depth is dependent upon exact
cancellation of signals.52 An exception to this situation is the field of circular arrays
where the symmetry can permit the compensation of mutual effects.53 Circular
arrays have proved attractive for steering nulls in omnidirectional type responses for
communication applications and this approach is mentioned in Chapter 12.
In addition to the null depth sensitivity to excitation errors, it is also necessary
to consider the effect of changes of frequency or bandwidth. Certain phase control
devices (such as the directional coupler) offer phase shifts which are independent
of frequency. This would result in a pattern in which the null directions changed
with frequency. Replacing the phase control devices with variable time delay
components will achieve directional responses whose nulls remain fixed in direction
with change of frequency.
Generally many null steering applications are limited to systems where it is pos-
sible to derive a feedback signal to verify the locations of the nulls. Such a system
may use an operator to apply the necessary corrections or a processor designed to
automatically steer nulls. Such systems are termed adaptive and are described in the
next section.
Array signal processing 399

13.7 Adaptive null steering

13.7.1 Introduction
A major disadvantage of the open-loop null steering arrays described above is that
the angular location and depth of the nulls is extremely sensitive to phase and
amplitude errors in the power splitting networks. In order to avoid this limitation,
an automatic control system is required which steers zeros in desired directions and
which compensates for any errrors due to tolerances in the array construction or
due to mutual coupling effects.
Data on the desired null directions is invariably obtained by processing some
form of measured co-variance matrix which defines the mean values of the cross-
products of the voltages received by the elements of the system. In the case of a
2-element system, the co-variance matrix M is:

*- [ 1 1 1
where the asterisk stands for the complex conjugate and an overbar indicates a time
averaged value, and Vu V2 = voltage received by element number 1, 2, respectively.
For two isotropic elements illuminated by a single narrow-band jammer at angle
d, the co-variance matrix reduces to

(Pt + Qn)
Kj Mn) , _ . , |
L P,exp(--#)
( l 3 6 6 )

where
Pj = mean jammer power at each element
Qn - receiver noise at each element (assumed uncorrelated)
0 = (27r<isin0)/A
In this case it is evident that the angle of arrival of the jamming signal can be
deduced immediately from the ratio of two terms of the co-variance matrix, since

(13.67)

In principle, the N xN co-variance matrix of an TV-element system can be manipu-


lated to provide angle of arrival data on up to (N— 1) jamming sources. This
involves analysing the incident signals into N eigenbeam components. This topic is
addressed in some detail later in this section, since it is an essential tool for under-
standing the mechanisms of adaptation. However, in the simpler adaptive arrays,
angle of arrival data is rarely evaluated explicitly, since it transpires that the opti-
mum element weights demanded by the widely-used maximum signal-to-noise
algorithm (described later) can be found simply by inverting the co-variance matrix.
The measured co-variance matrix is usually the result of averaging a number of
samples taken over a finite time duration. (To obtain decorrelated samples, the
interval between samples should be at least 1/fi, where B is the maximum band-
width of the system.) The measured values are thus contaminated by both receiver
400 Array signal processing
noise and by the effects of the noise modulation of the jammers. The number of
samples averaged should ideally be large, to provide adequate smoothing, but not so
large as to be unable to follow the angular movement of the jammers. Whilst the
number depends upon the detailed configuration of the adaptive array and upon
the environment in which it is to operate, a practical lower limit is about 2N
samples.
In the absence of interference, it is generally required that the radiation pattern
of the array should revert to a fixed quiescent pattern. For radar and fixed point-to-
point communication antennas, the quiescent pattern will usually be a single nar-
row beam with low sidelobes corresponding to perhaps a Chebyshev illumination
function. For mobile communication applications, where the directions of wanted
signals are unknown a priori, the quiescent pattern may be omnidirectional. When
interference occurs, the radiation pattern is required to adapt to present nulls in the
directions of the interference, but it is essential that the adaptive cirucits do not
attempt to cancel wanted signals. This is not much of a problem in pulsed radar
applications, since the wanted signal energy is generally very small compared to the
total jamming energy and high level signals from short-range clutter can be excluded
by time gating. In communication applications where both signals and jamming are
continuous, some means of distinguishing signals from jamming should be incorpor-
ated. In some cases, the jamming is assume to occupy a wider bandwidth than the
signals and the latter are excluded from the processing circuits by simple band-stop
filters. The same principle can be applied to certain types of spread-spectrum modu-
lation,49'54 since the signal modulation can be removed effectively once code lock
has been achieved. In cases where the signal modulation is not well defined and can-
not be distinguished from jamming, the simplest approach is to incorporate some
form of spatial filtering which prevents the adaptive circuits from responding over
certain defined angular sectors. Such additional constraints inevitably consume
some of the degrees of freedom of the processor, increase its complexity and
increase processing time.

13.7.2 Sidelobe cancellers

13.7.2.1 Single loop canceller: The simplest form of adaptive null steering
antenna is the 2-element sidelobe canceller. In this type of system, one element is
usually a high gain antenna, which may be an array or a reflector, and the other ele-
ment is virtually omnidirectional. Signals received by the omnidirectional element
are adjusted in amplitude and phase and used to cancel those received by a sidelobe
of the high gain element. (The maximum amplitude available from the omnidirec-
tional channel is restricted to prevent main beam cancellation,) Since the system
has only two basic antennas, only one interference signal can be cancelled
effectively because the combined antenna has only one independently steerable
null, although the high gain element may have many nulls in its radiation pattern.
Consider a simple array such as that shown in Fig. 13.51. The signal processor
consists of a conjugate multiplier, a filter, an amplifier with gain G and a signal
Array signal processing 401

weighting multiplier. The conjugate multiplier and the filter effectively correlates
the output of the omnidirectional element with the output of the array. In practice
conjugate multiplication is achieved by selecting the lower (i.e. difference) sideband
of a mixer.55

F ig. 13.51 Single -loop sidelobe canceller

The output of the array is


vo= vl-wv2 (13.68)
where Vx and V2 are the complex voltages produced in Ex and E2y respectively by
incoming interference with an angle of incidence 6 radians from the broadside
position.
Assuming a simple first-order filter, then the output z is related to the input U
by:
U = TZ + Z (13.69)
where r = filter time constant and z = dz/dt. Generally the bandwidth of the corre-
lation loop is small compared to that of the input signals, and therefore quantities
in the feedback path can be written in terms of their averaged values. Thus

u = vo\q= (13.70)
where V\ is the conjugate of V2 and an overbar denotes an average value. Referring
to Fig. 13.51, it follows that
W
z = (13.71)
402 Array signal processing

and combining eqns. 13.69,13.70 and 13.71 gives the differential equation

j;W + W(l/G + Wtf) = Vjq (13.72)


G
A general solution to the first-order differential eqn. 13.72 is given by

where Wo is the initial value of W at time t = 0.


The steady state solution is

Moo) = — ± » -i-i Q 3 74)

when G~l < |F 2 | 2 . If it is assumed that a wanted signal from the broadside position
has insignificant average power compared with the interference signal received from
some other angle, then the signal in the omnidirectional element is almost entirely
due to the interference. This is an important assumption because eqns. 13.73 and
13.74 make no distinction between signals received by a sidelobe or the main beam
of the high gain element. If there is sufficient gain in the processor loop, large
wanted signals received by the main beam of the high gain element could be
cancelled.
An interesting property of eqn. 13.73 is that the effective time constant for the
transient response is
r r

Thus the time required for the array to settle or converge is determined by the
interference power in the omnidirectional element, in addition to the loop para-
meters. Therefore the convergence is slow when a small interference signal is
received and fast when a large interference signal is received. If the interference
is very large the system time constant will be small and it is possible that the pro-
cessor will attempt to track the noise and fine structure of the interference signal.
If this happens the output signal will become noisy because of the rapidly varying
signal weighting function applied to signals from the omni-directional element. As
the incoming interference level is increased a point will be reached where the
system becomes unstable and the output oscillates. The rapid variations in output
signal caused by a small time constant occurring in eqn. 13.73 are known as loop
noise.56
In practice, the gain G must be chosen in conjunction with the maximum expec-
ted level of interference |F 2 | 2 to ensure that the loop bandwidth (== 1/2ITTC) is less
than 10% of the input RF bandwidth.
In order to reduce loop noise the time constant of the filter has to be increaseed
or the loop gain decreased, but this would result in even slower convergence when
weak interference signals are received. One solution to this problem is to insert an
amplitude limiting device at point A in the circuit of Fig. 13.51 which would
Array signal processing 403

maintain the value of \V2\2 constant for signals above a set threshold level. If the
device were an automatic gain control rather than a hard limiter the signal ampli-
tude could be controlled without distorting its phase characteristics. If automatic
gain control is used, terms in V2 can be replaced by K exp j<t>2, where K is a constant
and <j>2 is the phase of the signal V2. The convergence would then be determined by
G, r and K2, and would be independent of the interfering signal level.
Consider an interference signal incident at some angle Qt radians to the broadside
of the array shown in Fig. 13.51. For simplicity assume that the interference has
both a plane wavefront and is coherent over the period of time required for the
control loop to converge. In addition, further assume that no wanted signal or noise
is present. In this ideal situation

] (13,5,
V2 = G{d)VJ4)
where G^Of) and G2(6t) are the complex polar responses of the high gain and omni-
directional elements respectively at #,- with respect to an arbitrary reference
element.
VR = voltage induced in the reference element by the incoming interference signal,
2nd
(j) =
A
d = element spacing
X = interference signal wavelength
Inserting eqn. 13.75 into eqn. 13.74 gives

and the array output in the steady-state condition is


y(0t) = Gl(di)VR~W(oo)G2(6i)VRexpj<t> (13.77)
If the value for W(°°) given by eqn. 13.76 is substituted in eqn. 13.77, the resultant
output is zero and the array has produced a perfect null in its polar response at the
angle 0(.
In steering a null to the angle dt the array could possibly distort its main beam
pattern. The amount by which the main beam is distorted depends on several fac-
tors including:
(a) The location of the interference Of. If the interference is in either the main beam
or a near sidelobe the distortion is greater than if the null were required near an
existing one in the quiescent polar response.
(b) The polar responses Gx(8) and G2(6). In general the polar response of the high
gain element (Gi(6)) has a narrow beamwidth with, in a well designed element, a
low average sidelobe level. Therefore interference signals will not cause much distor-
tion of the array's main beam until they are within the main beam of the high gain
element.
404 Array signal processing

The example discussed is very idealised. In practice there is 'wanted signal' in the
high gain element and some residual in the second. Both elements will also have
some level of random noise. The effect of these additional signals is to: 56
(a) alter the loop convergence
(b) offset the null from the desired position
(c) soften the resulting null
The precise effect of these additional signals can be determined by including them
in the terms for Vx and V2 in eqn. 13.76.
An interesting consequence of the presence of random or receiver noise terms is
that the cancellation of the interference signal is limited by the noise power in the
omni-directional element. The cancellation of the interference cannot exceed the
interference signal-to-noise ratio in the omni-directional element, i.e.

^(MAX) ~ ~ ^ (13.78)

where
Pn = interference signal power received by the high-gain element
Pj0 = interference signal output power of the array
PI2 = interference signal power received by the omni-directional element
N — receiver noise power in the omni-directional element.
When this ratio is small, little cancellation occurs. This problem is not serious
because it is only significant when low level interference is received.

13.7.2.2 Bandwidth: The dispersive effects of the feed system and path length
mismatches limit the canceller's broadband performance. Long feeders between
elements and the canceller require careful matching in both dispersive and fixed
delays to avoid differential phase distortion. The spacing of the antennas can be a
further source of difficulty. When a large high gain element is used it may be
impossible to place the two antennas closer than several wavelengths apart. It is
possible that the subsequent delay introduced in one channel cannot be equalised
to that in the other for all directions. Generally, the frequency sensitivity of the
high gain element's polar response is the dominating factor in determining the
arrays broadband performance.

13.7.2.3 Multiple-loop sidelobe cancellers: The single cancellation loop sidelobe


canceller can only produce a single sidelobe null and cannot deal with interference
simultaneously received from more than one direction. In the presence of more
than one input the single loop tends to cancel the largest input to below the level of
the next largest. It actually equalises the products of the largest and next largest
pairs of voltages presented to the correlator after cancellation.
To cancel jamming from more than one direction requires as many loops as there
are jammers. Each loop is supplied from a separate omni-directional element with a
differing phase centre or from a separate element with a differing angular response
Array signal processing 405
(or both) . Individual loops do not in general act against individual jammers. The
weighting function for each loop modifies the composite jamming signals in that
loop and cancellation occurs when all the weighted signals are subtracted from the
primary high gain channel.
The conditions required to achieve cancellation of TV interfering sources can be
described by N simultaneous equations. If G0(6j) is the complex gain of the high
gain channel in the direction of jammer / and Gk(dj) is the gain of the fcth low-gain
channel in the direction of the same jammer, then a multi-loop system will achieve
perfect cancellation for weights Wx . .. Wk . .. WN given by:

WlGl(8l) + . .. WkGk(ex) + . .. WNGN(0x) = G0(fli)


WlG1(Q2) + . . . WkGk(62) + . . . WNGN(02) = G0(d2) (13.79)

.. . WkGk(6N) + .. . WNGN(6N) = Go(dN) )

The solution of this set of simultaneous linear equations can be expressed as the
ratio of two determinants. Thus for the M i loop:

Gi(P1)G2(61)...Go(0l)...GN(61)
Gi(O2)G2(02)...Go(e2)...GN(d2)

Gi(eN)G2(dN)...G0(6N) ... GN(6N)


Wh =• (13.80)

Gi(e2)G2(02)...Gk(02)...GN(82)

Gi(0N)G2(dN) ... Gk(6N)... GN(6N)

The equations become ill-conditioned if the determinant in the denominator of


eqn. 13.80 approaches zero. A simple case of two jammers and two loops will illus-
trate this condition. The required values of the two weights are:
G0(d1)G2(d2)~~G2(6l)G0(82) \
Gl(d1)G2(d2)~G2(dl)G1(d2)
and (13.81)
w = Go(O2)Gl(Ol)-Go(0l)Gl(02)
2
Gl(dl)G2(62)~G2(dl)G1(62)
It is evident that extremely large values will be required for Wx and W2 for those
combinations of angles of jammers where
Gi(0i)G2(O2) « G2(Pi)Gt(02) (13.82)
When this condition occurs, each loop produces a large output voltage but the two
voltages effectively cancel and the effective loop gain for the main channel is very
small. This results in poor cancellation and a slow response. An illustration of how
406 Array signal processing
the cancellation is affected by the arguments of the complex gains of the two auxil-
iary antennas is shown in Fig. 13.52, where two equal power interferers are
assumed with a variable angular spacing. All gains are held constant, but the cancel-
lation varies between zero (when eqn. 13.82 applies) and a high value determined
by the loop constants. Adding a third loop will reduce the region of poor cancel-
lation to a much smaller angular extent.

-80 r

-150
10 15 20 25 30 35 40
angle to moving interferes deg

F ig. 13.52 Cancella tion o f two interference inpu ts

auxilliary channels
main channel

-*«-jconjugate

adder

Fig. 13.53 Mult/loop side lobe canceller

A schematic for a multiple loop sidelobe canceller is shown in Fig. 13.53. Limi-
ters can be included in the conjugate branch of each auxiliary channel in an attempt
to provide a fast response.55 However, each limiter is captured by the strongest jam-
ming signal with the result that this signal is cancelled rapidly, but the loop gain to
Array signal processing 407

lower powered jammers is roughly proportional to their amplitudes and the


response to them is much slower. The same phenomenon of rapid response to high-
powered jammers but a slow response to low-powered jammers is observed in many
of the adaptive linear arrays described below. A partial solution in both cases57 is to
replace the omni-directional elements by a beam-forming array, such as a Butler
matrix, so that each loop tends to deal with one individual jammer, However, there
is still some interaction if any pair of jammers is spaced by less than one beam-
width of the matrix. A further refinement is to replace the Butler matrix by an
orthogonal beam-forming network of the type shown in Fig. 13.44, with the net-
work constants adjusted (by further feedback loops) so that the jamming signals at
the various output ports of the network are substantially decorrelated.48 Applica-
tion of this technique to adaptive arrays is discussed at the end of this Chapter.

13.7.2.4 Cancellation at IF: The principle of operation of the sidelobe canceller


is maintained if the cancellation circuits operate at IF rather than at RF, provided
that the phase and amplitude relationships of the various signal paths are preserved.
To achieve this, components of the receiver prior to the cancellation point must
have extremely flat passbands. Generally, identical front-end amplifiers and identi-
cal bandwidth determining filters are required and any long feeder paths must be
carefully matched in amplitude and phase characteristics over the frequency band.
Design of the correlators and limiting amplifiers needs some care due to the wide
dynamic range of interfering signals encountered and to the need to minimise AM
to PM conversion.

13 JJ Adaptive arrays
The concept of sidelobe cancelling can be extended to arrays with many elements,
each additional element allowing an extra null to be steered.55'58'59 In a situation in
which an iV-element array is receiving interference from /jammers, then ideally /
nulls would be directed at the jammers, leaving the remaining N — J— 1 nulls to be
adjusted so as to optimise the performance according to some predetermined
criteria. Of course, the number of jammers is rarely known in advance (and may
even exceed the number of elements), and the most widely adopted criterion is
simply that of optimising the ratio of signal power to interference plus receiver
noise power.
Consider a linear array of TV elements such as that shown in Fig. 13.38, which is
receiving a wanted signal from a direction 0. In the absence of any interference, the
signal has to compete only against receiver noise, which can be assumed to be un-
correlated from element to element with equal mean noise powers. It is well known
that in this circumstance, the signal-to-noise ratio is optimised by combining the
element voltages with uniform amplitude weights which have phase shifts which are
simply the conjugates of the signal phases at the elements. In the presence of inter-
ference from one or more directions, the net noise (interference plus receiver) at the
elements exhibits partial correlation and uniform weights are no longer optimum.
408 Array signal processing

The problem of optimising the signal-to-noise ratio can be solved by a variety of


methods, 55 ' 58 ' 59 but all methods essentially revert to the problem of solving a num-
ber of simultaneous equations, for which matrix operations are generally applic-
able.60 Adaptive arrays appear to be ideal candidates for digital techniques where, at
an early stage in the receiver, the element voltages are converted into digital words
representing amplitude and phase (or in-phase and quadrature components) and all
subsequent operations are performed by computer-type algorithms. A particular
problem with matrix operations, especially when a large number of array elements
is involved, is that the number of multiplications required is very large. This can
lead to substantial rounding-off errors in the computations, even where double-
precision arithmetic is employed. Furthermore, it is rarely possible to perform all
the multiplications within the basic Nyquist sampling rate of the incoming signals
(l/2# for a system bandwidth of B) and in some cases this makes the speed of res-
ponse a little slower than that achievable theoretically. An alternative implementa-
tion technique employing analogue or digital feedback techniques is often used.
This has many similarities with the multiloop sidelobe canceller and, since it senses
error signals and attempts to reduce these to zero, it is somewhat less prone to
rounding-off errors than the 'feed-forward' type of digital implementation.

13.7.3.1 Maximum signal-to-noise plus interference criterion: The derivation


given below is not confined to linear arrays. It is applicable to arrays where the
elements are located semi-randomly and where they may have differing beam
shapes.
Consider an jV-element array (e.g. Fig. 13.38). It is convenient to introduce a fic-
titious isotropic reference element, located inside or outside the main array, which
acts as the phase and amplitude reference for the voltages induced in the real ele-
ments. Suppose that the signal direction 6 is known and that the instantaneous
complex signal voltage produced in the isotropic reference element is a. Assume
that there are / narrow-band decorrelated jammers located at angles dx to dj, which
produce instantaneous voltages Ix to Ij in the reference element. Then the signal
voltage in the £th element is
A h = <xsk (13.83)
where
Sfe = Sk(0) = complex voltage gain of fcth element in direction 6 with
respect to the isotropic reference element.
similarly, the noise plus interference voltage at the kth element can be written as:
q =J
Ek = nk + I IqGkq (13.84)
q l
where ~
Gkq — Gh{Qq) = complex voltage gain of fcth element in direction Bq
with respect to the isotropic reference element
and
nh = receiver noise voltage in fcth element.
For the simple case of a linear array of N isotropic elements spaced d apart, it is
Array signal processing 409

convenient to place the fictitious reference element in the 'zero' position located at
—d from the first element. Expressions for sk and Gkq then become:
sk(6) = exp [(J2irkd sinO)l\] (13.85)
and
Gk(0q) = exp [Q'lnkd sindq)/\] (13.86)
When the element voltages are weighted by the factors Wx to WN and added
together, the mean power output due to signal is
I (Wksk) I (WX)\ (13.87)
whilst the mean power output due to interference plus receiver noise is

pn = r I (wkEk)Pi (W;E;)\ (13.88)


where the asterisk is the complex conjugate and an overbar indicates a time average.
The problem is thus to find the values of Wx to WN which maximise the ratio Ps/Pn.
Ps/Pn can be regarded as a function in the TV-dimensional space (or, more strictly,
in 2iV-dimensional space to allow for amplitude and phase) defined by the weights.
At its maximum, P8/Pn must be stationary for minor perturbations in any of the
weights. Consider a particular weight Wk and assume that a small change AWk pro-
duces changes APS and APni respectively, in the signal and noise plus interference
powers. Then ifP8/Pn is a stationary value:

^±^ £ (13.89)
Pn + APn Pn

from which it follows that

- ^ = 5? = H (13.90)
*Pn Pn
where H denotes the stationary value. To be more specific, assume that the weight
Wk is represented by an amplitude and by a phase term:
Wk = Rkexp(jtk) (13.91)
then the derivatives of Wk and of R/£ for amplitude and phase perturbations can be
written as:

whilst
^ = jWk and ^ = -jW*k (13.93)

Consider first a change in amplitude only. Differentiating Ps with respect toRk and
using the above results yields
410 Array signal processing

p=i

Similarly, differentiating the noise termP n gives

^f = f\wh I (H^FfcJS;) + H^j I (WpEtEp)] (13.95)

Taking the ratio of these last two equations and using the relationship for the
stationary value H yields
P=N P=N
I (WpSp)
l

^ P=l P=l

A similar relationship, obtained by considering variations of phase \pk alone, is


p=N p=JV
oca*\(WkSk)
»- .' ,.» ' " wtr^- r-' < 13 - 9 ')

It will be noted that the two equations above differ only by the signs in their
numerators and denominators and that the right-hand side of each term in a large
bracket is the conjugate of its left-hand side. Hence they can be combined to yield
an equation which has to be valid for all values of k:

(wpSp)
(WpE*kEp) (fc=l,2,...,JV) (13.98)
H
Now the term %(WpSp) is the same in all k equations above and, furthermore, the
value of Ps/Pn will not be affected if all the weights Wx to WN are multiplied by
any arbitrary constant. The above equation can therefore be reduced to

I (WpElEp) (13.99)
p=i
where C is any arbitrary constant. Applying this for all values of k enables the
weights to be related by the set of simultaneous equations:
W1E*1El + W2E\E2 + . . . 4- WNE\EN = Cs\
^ £ I + . . . + ^£,,= . (i3ioo)

The above result is often derived by matrix algebra which enables the set of simul-
taneous equations to be written in compact form as
Array signal processing 411

M-W = CS* (13.101)


where
si

S* = (13.102)

is a column vector defining the wanted signal direction, i.e. the 'steering vector',

W = (13.103)

is the colum vector defining the optimum weights and

EXE, EXE, . .. E\EN 'EX' [EXE2

= El
E*2E2 . E2EN (13.104)

E^Et E%E2 . .. E^EN


.Ek
is the co-variance matrix of the interference plus noise at the elements (i.e. it
should exclude signal/signal cross products). E is the vector of element voltages and
ET is its transpose (rows interchanged with columns).
The formal solution to the problem of maximising the signal to noise plus inter-
ference ratio can thus be written as
W = (13.105)
l
where MT is the inverse of the matrix M.
It should be noted that this solution does not depend upon a detailed prior
knowledge of jammer locations and power levels, nor is such information evaluated
explicitly. Theoretically, if the co-variance matrix can be measured accurately, then
it is possible to consider an open loop system of determining the weights which
simply solves the above matrix equation. Such a method is not easily implemented
(except for very small numbers of array elements), and a method which checks the
actual noise plus interference output power and applies feedback correction is often
preferred. Furthermore, the co-variance matrix can become ill-conditioned in cer-
tain circumstances (e.g. where there are closely spaced jammers) and, although its
determinant never becomes zero (due to the receiver noise components in the lead-
ing diagonal terms), a very large number of samples may be required to obtain an
accurate assessment.
An important advantage of the maximum signal-to-noise-plus-interference criter-
ion is that it automatically compensates for any errors in the array caused by mutual
412 Array signal processing
coupling effects or manufacturing tolerances. Such errors can be represented as
differences in the complex voltage gains of the elements as functions of beam angle
and, whilst these will modify the values of the various terms in the co-variance mat-
rix, the optimum weights are still given by eqn. 13.105.

13.7.3.2 Matrix inversion: As noted earlier, direct inversion of the co-variance


matrix is rarely used in the simpler practical adaptive arrays. However, it is fre-
quently required in calculating the theoretical performance of such arrays.
Denoting by mjk the element in the /th row and kth column of the co-variance
matrix
mn mn . . . mXN

M = m21 m22 (13.106)

m
mm NN

then the inverse matrix is:

M \M\ \M\
An A22 Am
l
M~ = \M\ \M\ \M\ (13.107)

AW Aw
\M\ \M\

where \M\ is the determinant of M and AJk is the cofactor of m-ih (i.e. the reduced
determinant obtained by deleting the /th row and the &th column, with a sign given
by (—l) J+fe ). The formal procedure for finding the inverse is thus to replace each
element mjk by the quotient of its corresponding cofactor and the determinant \M\
and then to interchange rows and columns.
It will be appreciated that, if the determinants are evaluated by multiplying out
the various terms in accordance with Cramer's rule, excessive numbers of multipli-
cations are required and the rounding-off errors can be substantial. It is more usual
to solve the simultaneous equations (eqn. 13.100) directly by elimination tech-
niques which successively subtract multiples of one row from another so as to con-
vert the original equations from the form
Wxmn + W2m12 = Cs\
W2m22 WNm2N = Cs*2 (13.108)

Wxmm 4- W2mm WNmNN = Cs%


into the form
A rray signal processing 413

W2m12 WNmm = Cs\


W2x22 (13.109)

where all the terms below the leading diagonal have been reduced to zero. (The x
and y above represent the modified coefficients for the other terms.) The equations
can then be solved for the weights by back substitution, starting with the last equa-
tion for Wjsf. This method reduces the number of multiplications required, but some
care is required in selecting the order in which the elimination proceeds in order to
minimise rounding-off errors.61 Even so, double precision arithmetic is generally
required in the computations.
For large arrays, iterative methods of solving the simultaneous equations are
sometimes used.62 Although these generally require more computation time than
the elimination methods, they are more economical in the use of computer memory.
Furthermore, they preserve the original matrix (unlike the elimination method
where the original is destroyed as the computation proceeds) as a means of testing
the latest iterative for the weight vector and the rounding-off errors are usually less
severe.
Discussion on the methods used in practical arrays is deferred until after some
further properties of the co-variance matrix have been described.

IS. 7.3.3 Steering vectors incorporating amplitude weighting: For a linear array,
the simplest steering vector for a wanted signal direction 0 is

exp[(--;2m/sin0)/A]
* _ exp[(—j4nd sin0)/X]
(13.110)

exp[(—j2Nnd sint

In the absence of interference, the noise co-variance matrix is simply

"Qn 0 ... 0"


0 Qn ... 0
M = (13.111)

0 0 Qn
where Qn is the receiver noise power per element. Applying this to equation 13.100,
it can be seen immediately that the optimum weights in the absence of interference
are
C
w = —s* (13.112)
Qn
414 Array signal processing
Thus, as expected, the system reverts to uniform amplitude weights in the quiescent
condition.
In many applications, array designers are willing to sacrifice perhaps 1 or 2 dB in
signal-to-noise ratio in the quiescent condition in order to obtain lower sidelobes.
This can be incorporated readily by including the quiescent design weights b x to b^,
which control the quiescent sidelobe level in the desired manner, into the steering
vector. Thus the new steering vector is taken as:
exp [(-/2m/ sin 0)/X]
[(~-j4nd s
S* = (13.113)

J)N exp [(-jlNnd sin 0)/X]_


Clearly, eqn. 13.112 still applies and the adapted pattern in the absence of interfer-
ence is the desired low sidelobe design.
In the presence of interference, the shape of the adapted pattern depends to
some extent upon the design weights incorporated into the steering vector, but it is
still optimised in the sense that nulls are directed towards the jammers. Fig. 13.54
shows some theoretical adapted patterns for a 4-element linear array where, for sim-
plicity, the signal direction is taken as broadside and two narrow-band jammers are
assumed at angles 6X and 62. In this case, the theoretical noise plus interference co-
variance matrix is

(Qn+Pl+P2)
(Qn +

(Qn P X + P 2)
(13.114)
=
where Pu P2 mean interference power per element due to jammer number 1, 2
and 0x = (2nd sin #i)/A and 0 2 = {^d sin 02)A-
The theoretical adapted patterns are obtained by evaluating the inverse of ilf and
multiplying by the appropriate steering vector. The first diagram in Fig. 13.54 cor-
responds to uniform weighting, i.e.

(13.115)

whilst the second diagram shows the effect of incorporating binomial amplitude
weighting into the steering vector:
Array signal processing 415

10 dB

-30
-180

^ = 10 dB
416 Array signal processing

-20

-30
-180° -90* 90° 180°
2tTd sine

(c)

Fig. 13.54 Adapted radiation patterns of a four element array with two jammers
a Uniform amplitude steering vector S*T = [ 1 , 1 , 1 , 1 ]
b Binomially-weighted steering vector S*T = [ 1 , 2, 2,1 ]
c Omnidirectional steering vector S*T = [ 1 , 0, 0, 0]
quiescent pattern
adapted pattern

Adapted weights
(a) Uniform (b) Binomial (c) Omnidirectional
0.4056+/0.0210 0.3837 +/0.0852 0.7535+/0
W2 0.5627+/0.1359 0.5778+/0.1080 0.0433 -/0.0193
0.5627-/0.1359 0.5778-/0.1080 0.4886 -/0.2777
0.4056-/0.0210 0.3837 -/0.0852 -0.2349 +/0.2426

o*
(13.116)

The third diagram is relevant to the case of a mobile communication system in


which the desired signal direction is unknown. An omnidirectional quiescent
pattern is therefore required and the steering vector corresponds to exciting only
one element of the array, i.e.
Array signal processing 417
V
0
s* = 0 (13.117)

The nulls in these example patterns are rather broad, since there are only four ele-
ments in the array. With a greater number of elements, much sharper nulls are
obtained.

13.7.3.4 Additional main beam constraints In deriving the maximum signal to


noise plus interference condition
W = CM~X'S* (13.118)
it was stated that C could be an arbitrary constant. Whilst this is theoretically cor-
rect, since it is only the relative values of the weights which affect the signal to
noise plus interference ratio, there may be some advantage in allowing C to differ
for differing measured environments. For example if C were chosen as unity, then,
as the average jamming power increased, the average weight would have to decrease,
i.e. the system would tend to AGC on the average jammer power and the absolute
level of output signal would be reduced. This could be undesirable if the next stage
in the receiver were, say, a communications demodulator requiring a substantially
constant input signal level. In this case, it would be desirable to introduce the con-
straint

S*TW = £ S$WP = 1 (13.119)


P*I
which maintains the signal output at a fixed level.
In the case of a mobile communications antenna where the desired quiescent
pattern is omni-directional, a simple method of reducing the dynamic range of
weight variations is to clamp one element, usually an end element.63
In other applications it may be better to adopt the 'zero-order' constraint:
p*N
I \WP\2 = N (13.120)

This constraint ensures that the RMS weight value is unity and maintains that part
of the output due to receiver noise alone at a constant value. Since large jammers
are usually cancelled to well below receiver noise, this constraint is fairly effective
in maintaining constant false alarm rate properties for radar receivers.
The above constraints are essentially single point constraints and do not affect
the operation of the maximum signal-to-noise-plus-interference criterion, apart
from the change in the scale factor C. However, they may become inadequate if the
direction of wanted signals is now known precisely. In radar applications, for exam-
ple, the measured co-variance matrix is often allowed to include wanted signal
418 Array signal processing
terms, because they are difficult to exclude and it is argued that their average
energy is small compared to the interference and receiver noise energy. In some
modern high-duty-cycle pulse Doppler or pulse-compression radars, the signal
energy may approach the level of receiver noise energy and in this case the adaptive
circuits will begin to cancel the signal.
This situation is illustrated in Fig. 13.55, which shows the main beam of the
quiescent pattern, which is applicable when the signal energy is much smaller than
the receiver noise energy, and the adapted response (labelled 'zero-order con-
straint') as a function of signal direction for a signal-to-noise ratio of OdB. The
effective beam width is reduced to about one tenth of the nominal.

signal bearing error


0

/ f l \
S 10 I \ Vi-^^^-quiescent beam ( l o w S/N)

-first-order constraint
20

zero-order constraint

30

Fig. 13.55 Signal suppression at 'high' S/N ratios


S/N = 0dB

Additional constraints can be introduced to overcome this deficiency.54 One


technique is to control the first one or two derivatives of the pattern in the wanted
signal direction. For example, if it is desired to have zero slope in the adapted
pattern in the wanted signal direction, then, since the adapted pattern is

G(6) = I Wpexp(jp<j>) (13.121)


p=sl
where
<t> = 2ndsind/X
the first derivative of the pattern with respect to sin# becomes
8G(0) %N
—^ = (}2nd/X) I pWpexp(Jpct>) (13.122)
o sin d p* i
and this must be constrained to be zero. This constraint consumes one degree of
freedom in the array, since, once the first TV— 1 weights have been found, theMh
weight follows from:
A rray signal processing 419
_j p=JV-l
W
N = -jj~ I pWpQxpj(p-N)(p (13.123)

Expressions for the mean signal power and for the mean interference plus noise
power corresponding to eqns. 13.87 and 13.88 can then be set up by making the
substitutions

and } (13.124)

and restricting the summations to N — 1 terms. The derivation of the optimum


signal-to-noise-plus-interference can then proceed along the same lines as before,
yielding N — 1 equations which are similar in form to, but with slightly more com-
plicated coefficients than, eqn. 13.100. These can be solved to find the first / / — 1
weights, with the M h weight being given by eqn. 13.123.
The beneficial effect of introducing a zero-slope constraint is shown in Fig.
13.55 by the curve labelled 'first order constraint'. Techniques for introducing
higher order constraints are given in Reference 54.

13.7.3.5 Retrodirective beams: The operation of an adaptive array can be


envisaged as adding a number of retrodirective beams to the quiescent pattern. Con-
sider a 4-element array with a single jammer at angle d1. The simultaneous equations
relating the optimum weights W to the steering vector S* are:
= Cs\

Wx{Pie'm^ 4 W2(P1e~J^) 4- W3(Qn + P{) 4 W^P^) = Cs3

These can be re-arranged into the form

Wk = ^-Bexp(-jk<t>i) (it = 1 , 2 , 3 , 4 ) (13.126)


where
p
B = (W e*^1 4 W e*2^1 4 W e*3^1 4 W e ^ 1 ) (13 127)
Qn l 2

can be regarded as a constant multiplier. Eqn. 13.126 indicates that each adapted
weight Wk is equal to the quiescent weight minus a uniform amplitude weight (\B\)
phased to direct a beam at the jammer. Eqn. 13.127 indicates that the amplitude of
the retrodirective beam is proportional to the jammer power Px and to the ampli-
tude of the final adapted beam at <plt The absolute phase of the retrodirective beam
is such as to achieve substantial cancellation of the jammer.
420 Array signal processing

The concept above can be extended to cover / jammers, even if / exceeds the
number of array elements N. Thus the adapted weight vector can be written as

(13.128)
Qn i=l
where
(13.129)

is the weight component corresponding to the retrodirective beam pointing at the


zth jammer.
An example of analysing the adapted pattern of the two jammer situation of Fig.
13.5% into the quiescent pattern plus two retrodirective beams, each with uniform
amplitude weights, is shown in Fig. 13.56.

jammer 1 jammer 2

lOdB

-ieo #

Fig. 13.56 Analysis of adapted pattern as quiescent pattern and two non-orthogonal retro-
directive beams
Quiescent pattern
Retrodirective beam 1
Retrodirective beam 2
Adapted pattern

Whilst the above analysis gives a simple physical picture of the operation of an
adaptive array, it does not readily reveal the amplitudes of the various 'uniformly
illuminated' retrodirective beams (until the actual values of the adapted weights
Array signal processing 421
have been computed). The amplitude of each retrodirective beam depends in a com-
plicated way upon all the jammer power levels and directions, and the various
beams are generally not orthogonal to each other.
Further insight into the operation of the array can be obtained by analysing the
retrodirective beams as a weighted set of particular orthonormal beams known as
the eigenbeams, as discussed below.

13.7.3.6 Eigenbeams: Any radiation pattern can be analysed as the weighted


summation of a complete set of orthogonal beams. For example, Butler matrices
and radiation pattern synthesis techniques usually employ a set of similar-shaped
(sinAfr/sinx) orthogonal beams, spaced \/D apart.
Eigenbeams are a particular set of orthogonal beams which depend upon the
jammer powers and directions, but do not depend in any way upon the steering
vector S*. The weights XP to XJV) required to set up the /th eigenbeam define the
eigenvector X^\ Associated with each eigenbeam is an eigenvalue A,- which is a
function of the jammer powers and directions. For the simple case of a single jam-
mer it will be shown that there is only one unique eigenvalue, given by
Xi = Qn+NP, (13.130)
In this case, there are JV— 1 non-unique eigenvalues equal to the receiver noise
power per element, i.e.
X2 = X3 = .. . = XN = Qn (13.131)
For / narrow-band decorrelated jammers there are / unique eigenvalues, which
depend upon the jammer powers and directions, and//—/non-unique eigenvalues
equal to the receiver noise. The eigenbeams corresponding to the / unique eigen-
values generally have differing shapes. (If the jammers are widely spaced, i.e. by
several basic beamwidths, the shapes become similar and the eigenvalues are roughly
proportional to the individual jammer powers. Each eigenbeam peak then points
approximately at a different jammer.)
Strictly, there are an infinite number of sets of eigenbeams corresponding to the
non-unique eigenvalues, but all these 'non-unique' eigenbeams have zeros in the
directions of all the jammers. (Many computer programs which compute eigenvec-
tors and eigenbeams 'invent' a set of eigenbeams for the non-unique eigenvalues as a
result of rounding-off errors in the computations.)
Mathematically, eigenbeams are normalised to form an orthonormal set by
requiring that, for all possible combinations of beam numbers (z) and (/):

JW> i :7 03-132)
(/*/)
The first condition above simply normalises the gains so that the output power due
to receiver noise alone is Qn for all the eigenbeams. The second condition (i%j) is
the mathematical criterion for orthogonality, but it is worth noting that it produces
strictly orthogonal beams only if the radiation pattern of each element of the array
has a special form. If
422 Array signal processing
E(d) = complex voltage gain pattern of each array element (13.133)
then the overall radiation pattern is

G(6) = £(0)2, Xkexp(kj(p) (13.134)


where
<j> = (2nd sin 6) /X
The condition for beams (z) and (/) to be orthogonal is that their conjugate cross
product, averaged over all angles of real space, shall be zero, i.e.
7T/2

J Gi(pyGj(d)* dd = 0 (13.135)
-7T/2

Substituting eqn. 13.134 and changing the variable of integration from 0 to 0 yields

J0 (13.136)

Now it is a well-known property of Fourier series that

— I exp[j(k — p)<t>]d(j) = I (13.137)


27r
2n J n fir v- r>\

and hence eqn. 13.136 reduces to eqn. 13.132 if

fVcosJ (\sm6\<\l2d)
\E(B)\ - (13.138)
I 0 (|sin(9i>A/2d)

The last equation represents Wasylkiwskyj's Ideal pattern,64 which has a eos0
power pattern, but is zero beyond |sin 6 \ = \/2d. (This corresponds to the scanning
angle at which grating lobes just begin to appear.) Much of the theoretical treatment
of adaptive arrays implicitly assumes this ideal element pattern, but practical arrays
rarely achieve it due to mutual coupling effects. However, as noted earlier, mutual
coupling effects are automatically compensated for by the maximum signal-to-
noise-plus-interference criterion and the idealised conditions assumed enable the
basic mechanisms of adaptation to be understood.
Returning to the definition of eigenvalues, if the output noise plus interference
power Pn is regarded as a function in the 2Af-dimensional space defined by the
weights, then each eigenvalue \t corresponds to a maximum value of Pn.
Eqn. 13.88 gives the expression for Pni but in order to take account of the
normalisation given by eqn. 13.132, the quantity to be maximised is
A rray signal processing 423

fe-iV p=M
I (xkEk) I
k=l
(X;E*P)
P — (13.139)
kN \
k
I (xkxt)\j
Maximising the ratio (interference-plus-noise)/(noise) in this way enables each
weight to be considered as a completely independent variable and the interrelation-
ship inherent in eqn. 13.132 to be accounted for later.
The maximisation of Pr can proceed along the same lines as before, differentiat-
ing with respect to amplitude Rk and phase \pk of each weight Xk. (cf. eqn. 13.91
et seq.). The details are omitted here, but the final result is another set of simul-
taneous equtions:

(13.140)
1
where Xs is a maximum of Pr (i.e. an eigenvalue).
Alternatively, these equations can be expressed in the deceptively simple matrix
form:
\SX = M-X (13.141)
Written out in full, the equations are

t = XlE\El + X2E\E2 Xjs[EtEN


(13.142)

Essentially, they form an overdetermined set of N equations in the N — 1 variables


X2/Xi to XN/XX and the set is consistent only if the determinant

(E\E2)
(E2EO (E2E2 ~ (E*2EN) = 0 (13.143)

(E*NE2)

where Xs is an eigenvalue. In matrix form, this characteristic equation is


\{M-\J)\ = 0 (13.144)
where / is the identity or unit matrix.
The eigenvalues are thus the roots of an iVth degree polynomial in Xs. They can
be found by any standard polynomial root finding procedure61 and the eigenvectors
determined by substituting back into eqn. 13.142. Except for low-degree poly-
nomials, numerical root-finding procedures are inevitably iterative and the number
of iterations required cannot be predicted in advance. The number of iterations is
424 Array signal processing

usually large when there are closely spaced eigenvalues. For this reason, eigenbeam
analysis is not used in simple practical adaptive arrays, but is a useful tool for theo-
retical analysis and for more complicated arrays where a detailed analysis of the
jamming environment is required. Several iterative numerical procedures exist
which determine eigenvalues and eigenvectors simultaneously.62
Since the eigenvalues correspond to power ratios they are always real and posi-
tive. Mathematically, this property is associated with the fact that the co-variance
matrix is Hermitian, i.e. terms in complementary positions on each side of the lead-
ing diagonal are complex conjugates.
Since Xs appears only in the diagonal terms of the determinant (eqn. 13.143), it
follows that when the determinant is expanded as a polynomial:
coefficient of Xf = ( - 1 ) * (13.145)
and fe=N
coefficient of xf" 1 = (~ l ) ^ " 1 £ E%Ek (13.146)

(The coefficients of lower powers of Xs are more complicated). The negative of the
ratio of these two coefficients gives the sum of the roots for any polynomial and
hence S=N U=N
I Xs = I E*kEk (13.147)

Thus the sum of the eigenvalues is the total interference plus noise power incident
on the array elements.
Having determined the eigenvalues, each corresponding eigenvector X^l) can be
obtained by substituting the value of Xf back into the simultaneous eqns. 13.142
and solving these for the various weight components X\1^ to X$. Since these N
equations are essentially an overdetermined set in the iV— 1 variables X2/Xi to
XJV/XJ, any N — 1 of the equations can be solved for these weight ratios and the
actual weight values found from the normalising condition
fe = iV
I \Xk\2 = 1 (13.148)

(The problems of the non-unique eigenvectors associated with the non-unique


eigenvalues equal to receiver noise is discussed later.)
Assume that a complete set of eigenvectors has been determined. Then a beam-
forming network or matrix transformation, such as that shown in Fig. 13.57, can be
constructed which converts the input signal voltages Ex to EN into the eigenbeam
voltages Ui to UN. Thus for each eigenvalue \t the eigenbeam voltage is
fc=JV

Ut = I 4 % (13.149)
fe = i
The complete set of eigenbeam voltages can thus be represented by a vector
U = LE (13.150)
where
Array signal processing 425

\f
• /
E

w2

1
I
add

T r. ,-T.
2 (W k E k ) = W E = JE W

(a)

Fig. 13.57 Weighted linear array (a) and its eigenbeam equivalent (b)
426 Array signal processing

L = (13.151)

is the N by N matrix defining all the eigenvectors.


Since the basic optimisation problems involve the co-variance matrix M of the
element voltages E, it will be evident that in the transformed 'eigenbeam space'
interest should now be centred on the co-variance matrix of the eigenbeam voltages
U. Defining this matrix as

WN (13.152)

u*Nux u%u2 ...


then the general term in the fth row and ;th column is
p=N
(13.153)

Recalling that each eigenvalue is related to its eigenvector by the simultaneous


equations:
(13.154)
then k=N
U*Uj = \S (13.155)

Now for / = / the eigenvectors have been normalised by eqn. 13.148 and hence
UfUi = W=h (13.156)
Hence the sum of the eigenbeam powers is equal to the sum of the eigenvalues,
which has already been shown to be the total interference plus noise power incident
on the array. It therefore follows that
UfUj = 0 (13.157)
and that the eigenvectors form an orthonormal set, satisfying eqn. 13.132. (If the
eigenvectors were not orthogonal, it would be possible to form some combination
of eigenbeam voltages which increased the total power.)
Mathematically, L is the transpose of the 'modal matrix'. Since L is an orthogo-
nal matrix, its inverse is equal to its conjugate transpose:
L1 - Z* T (13.158)
and it transforms the co-variance matrix M of the element voltages E into a diagonal
Array signal processing 427

form of co-variance matrix \x of the eigenbeam voltages U. Thus:


M= (UW) = JEFfiLEf = I*(^*£T)IT (13.159)
or
Kx 0 ... 0"
0 X2 . . . 0
M = L*MLT = 1*3 (13.160)

0 0

For adaptive arrays, the advantage of an eigenbeam analysis is that the co-variance
matrix i$, reduced to the above simple diagonal form and that the various eigen-
beam voltages are decorrelated. This topic is addressed in more detail later, after a
few examples of eigenvalues and eigenbeams have been discussed.

13.7.3.7 Eigenvalues produced by a single narrow-band jammer: Consider a 4-


element linear array illuminated by one narrow-band jammer at angle Qx which pro-
duces a power Px at each element. The characteristic equation 13.143 is:

=0

(13.161)
where 4>l = (2n sin 0^/X.
The determinant can be reduced by the usual rules for operating upon determin-
ants. First, multiply the first column by e]k<t>l and subtract from the (k + l)th
column (k = 1,2,3). This gives

\) (K-Qn)^ (\-Qn)ei2^ (X, - Qn)e>


(Pie^) (fin-X.) 0 0
=0
(/V'2*') 0 (Qn-\) 0
(Pie~m>) 0 0 (Qn-K)
(13.162)
;fe
Next, multiply each (k + l)th row by e *' and add to the first row, to yield:

0 0 0
- o o = 0 (13.163)
0 (Qn-\) 0
0 0
428 Array signal processing
The determinant can then be expanded to give the characteristic equation as:
(Qn + 4Pl-\s)(Qn-K)3 = 0 (13.164)
which indicates that there is one unique eigenvalue
*i = Qn + Wi (13.165)
and three equal roots or non-unique eigenvalues
X2 = X3 = X4 = Qn (13.166)
The above reasoning can clearly be extended to an iV-element array with a single
jammer to show that the unique eigenvalue is
Xi = Qn + NPt (13.167)
The unique eigenvalue corresponds to the maximum interference power which can
be abstracted (NPX) plus receiver noise power corresponding to one element. The
non-unique eigenvalues each correspond to receiver noise for one element, and
hence the sum of the eigenvalues represents the total interference plus noise power
of the system.
Formally, each eigenvector XS1* can be obtained by substituting the appropriate
eigenvalue \ into the simultaneous equations and then solving these equations for
the various weight components X*p to X$.
However, in the present example it is easier to proceed by writing eqn. 13.142 in
the form
XXPX
X2PX

XxPxe-j^ + X2Pxe-j2^ 4- X^e"^ + XJ>X = (Xs - Qn)X4 j


(13.68)
which it is immediately evident that, for the unique eigenvector, the weight
ratios are given by
Xx = X2eJ*i =X3ej2^ =X4em> (13.169)
The normalising condition (eqn. 13.132) then gives |X^| = 0.5, so that the unique
eigenvector can be written, noting that the choice of absolute phase is arbitrary, as:

(13.170)

which represents uniform amplitude weights, phased to direct a beam at the jam-
mer. The radiation pattern of the unique eigenbeam at any angle 6 is thus
Array signal processing 429

G(d) = i Xke*« = 0.5 £ = 0.5


sin [(0 —00/2 J
i i

xcxp/[5(0-00/2] (13.171)
which is, of course, the usual (siniVx/sinjc) shape of a uniformly illuminated beam.
The phase angle of this beam is a result of having chosen the hypothetical phase
reference element in the 'zero element' position of the array. The peak amplitude
of the beam is 2, corresponding to a power gain of 4, equal to the number of
elements.
For the three non-unique eigenvalues Xs = Qn and all the equations 13.168
reduce to

= 0 (13.172)

which indicates that the corresponding eigenbeams must all have a zero at 0 = 0X.
Apart from the orthonormalising constraint (eqn. 13.132), eqn. 13.172 is the only
other constraint on these beams and it is clear that their eigenvectors are indeter-
minate. The orthonormal set can be completed by 'inventing' (sinNx/sinx) beams
of the same shape as the unique eigenbeam, spaced in 0 by n/2 as in a 4-element
Butler matrix. Thus the non-unique eigenvectors could be chosen as:

0 5 e"1 0" -
05
(1 = 2,3,4) (13.173)

05

However, it must be emphasised that this choice is arbitrary. For example, it can be
readily checked that the three weight vectors:
1 _,„ Q

0 -O.Se"' 301

0 0.5 e"

(13.174)
satisfy eqn. 13.168, are mutually orthogonal and are orthogonal to the unique
eigenvector (eqn. 13.170). Thus they, and many other possibilities, could be con-
strued as completing the set of orthonormal eigenbeams.
430 Array signal processing
In summary, there is only one unique eigenvector for a single jammer illuminat-
ing an TV-element linear array, given by

1 _,*

(13.175)

The other eigenvectors are indeterminate, but all non-unique eigenbeams have a
zero in the jammer direction. Thus, in theory, this direction can be determined
exactly. In practice, it will be disturbed by noise in the measured co-variance matrix
due to the finite, rather than infinite, measurement duration.

13.7.3.8 Eigenvalues produced by several narrow-band jammers: In any eigen-


beam analysis of a measured co-variance matrix, it is usually necessary to solve the
Mh degree polynomial (eqn. 13.144) for the eigenvalues. As noted earlier, many
computer programs will 'invent' eigenbeams for the non-unique eigenvalues as a
result of very small differences in these eigenvalues caused by rounding-off errors.
In theoretical analyses where the number of jammers / is less than N, a reduced
form of characteristic equation can be derived which is a polynomial of degree / .
The N equations which must be satisfied by the eigenvalues \$ are

= (K~Qn)X\

+ .. . + X = (ks-Qn)X2
N

(13.176)
where Pt and fa refer to the zth jammer and the summations extend from i = 1 to / .
It is convenient to introduce the function
p=N
= e
-4(0) 2- (13.177)

which is simply the pattern of a uniform amplitude constant phase excitation, with
phase referred to the first element. Note that A(— <j>) =A*(<f>). Now define
p=N
(13.178)
P=I
Array signal processing 431

so that the various Bt represent the complex amplitude of the as yet undetermined
eigenbeams at the various jammer directions 0,-. If each kth row (k = 1 to N) of the
equations is multiplied by e^k~1^1 and all the equations are then added together,
the result can be written as

Similarly, if each/rth row is multiplied by ej(fe~1)(^2and the equations added:

£ PiBtAifa-cpi) = (Ks-Qn)B2 (13.180)


1=1
Proceeding in this way,/ simultaneous equations in the variables Bx to Bj can be
derived. Noting that ^4(0) = N, the equations can be written out in full as:
BXPXN +B2P2A((t>x-4 ~~ 0j) = (Xs - Qn)Bl
B^Aifa-tJ+B^N . . . + BjPjA(<t>2 - 0j) = (Xs - 2 n ) £ 2

- 00 -02) (Xs - Qn)Bj


(13.181)
They form an over-determined set o f / equations in the / — I variables/2//i to
Jj/J\. The set can only be consistent if the determinant

i^ ( 0 2 - 0 0 = 0

(13.182)
Thus for theoretical problems where the jammer powers and directions are speci-
fied, the above reduced form of characteristic equation can be used to find the
unique eigenvalues. Note that these eigenvalues depend upon the jammer powers
and the differences in their bearings, rather than upon the absolute bearings. (The
N—J non-unique eigenvalues are given by Xs = Qn, as before.) Having determined
the unique eigenvalues, the corresponding eigenvectors can be found by substituting
back into the original equations 13.176.
Applying eqn. 13.182 to the case of two jammers, the determinant can be
expanded, giving the quadratic equation
(NPi + Qn- \WP2 + Qn - \) = PlP2\A(4>2 - 0OI2 (13.183)
2
The quantity l^4(0 2 ~0i)l is simply the power pattern at angle 4>2 — <i>l for a uni-
formly illuminated array and is reasonably small for jammer separations exceeding
one basic beamwidth of the array. In that case, the two unique eigenvalues are
432 Array signal processing

and (13.184)

The relations are exact when the jammer separation is a multiple of one beamwidth.

CQ

90° 180°
relative jammer spacing , I ^ - ^ I

Fig. 13.58 Eigenvalues for two jammers (Px/Qn = 20dB: P2/Qn = WdB)
4 element array
8 element array

Fig. 13.58 shows how the eigenvalues vary as a function of jammer separation. For
small separations, there is one large eigenvalue, approximately Qn + N{PX + P2), and
the second eigenvalue is just greater than receiver noise Qn. At large separations, the
lower eigenvalue is more affected by the exact jammer separation than is the higher
eigenvalue.
Array signal processing 433

beami

beam 2

beam 3

Fig. 13.59 (a) Orthonormal beam-former

jammer 1

30r

-180
2TTCJ
sin 9

Fig. 13.59 (b) Power observed with uniformly illuminated scanning beam {beam port 1)
Jammer 1 alone
Jammer 2 alone
Observed resultant
434 Array signal processing
Some physical appreciation of eigenbeams and eigenvalues can be obtained with
the help of the orthonormal beam-forming network of Fig. 13.44. Consider a
4-element array feeding such a network and assume that there are two jammers
present, as in Fig. 13.59. Suppose that a power meter, with adequate smoothing to
ensure it measures true mean values, is connected to beam port 1. Assume that the
first row of power couplers in the network has been set to the values 1/2,1/3 and
1/4, corresponding to uniform amplitude weighting, and that the first row of phase
shifters are adjusted in sympathy, with values \p, 2$ and 31^, so as to slowly scan
the beam throughout the full range of angles. The measured mean power output
will be as shown in the solid curve of Fig. 13.59.
At this stage, it will be apparent that there is a least one jamming source present,
but the response corresponding to the lower-powered jammer may be mistaken for
a sidelobe. Suppose that the phase shifters are returned to the peak power position
and that the various power couplers and phase shifters in the first row are then
gently perturbed in turn so as to attempt to increase the power meter reading.
Eventually, after several iterations, a true maximum will be found. This is the
largest eigenvalue and the weights corresponding to the coupler and phase shifter
settings in the first row form the eigenvector. The radiation pattern of beam port 1
is the corresponding eigenbeam.
The power meter can now be transferred to beam port 2 and a second scanning
beam set up with the couplers and phase shifters in the second row. For each scan
angle, the couplers and phase shifters should be set so as to produce maximum gain.
(The maximum realisable gain will generally be less than N owing to the spatial
filtering effect of the first eigenbeam.) These settings will involve somewhat messy
computations, since they depend upon the values set into the first row. Having
completed the scan, the settings can be returned to the greatest measured peak and
the effect of minor perturbations in the settings investigated. Since only two jam-
mers have been assumed, no further adjustment of the settings will be necessary.
(Had a third jammer been present, further adjustment would have been required.)
The output of the second beam is the second eigenvalue and the element weights
corresponding to the combined settings of the first and second row form the second
eigenvector.
If the power meter is now transferred to the third beam port and the procedure
followed to set up a third scanning beam, it will be found that the third beam out-
put power is constant (at receiver noise level) and is not affected by the settings of
the third row of controls. The first two eigenbeams have abstracted all the jamming
power incident on the array and the remaining two non-unique eigenbeams are in-
determinate. Any setting of the third row of controls produces two orthogonal
beams at ports 3 and 4, which have zeros in the jammer directions and which com-
plete the set of eigenbeams.
In theory, the network described above could be used to determine all the jam-
mer bearings, since the non-unique eigenbeams have zeros at all jammer bearings.
The third row of controls, which forms the only degree of freedom left in the
example, could then be set to produce a maximum realisable gain beam in a given
A nay signal processing 435

signal direction. (In a more general case N — J—l degrees of freedom would be
available to optimise signal reception.) In practice, such an open loop method of
control would be unsatisfactory because of difficulties in measuring mean powers
with sufficient accuracy and because of the interdependence of the various control
elements. However, the eigenbeam network, even when set up only approximately,
is a very useful pre-processor to put in front of a simple adaptive processor using
the maximum signal-to-noise-plus-interference algorithm, since it improves the over-
all convergence time.

13.3.7.9 Eigenbeam analysis of the maximum signal-to-interference-plus-noise


criterion: Consider the array shown in Fig. 13.57a. The basic optimisation problem
is: 'Given the interference pluse noise voltages Ex.. . EN and a desired signal direc-
tion (or, more specifically, a signal steering vector 5*), what values of weights
Wi... WN&re required to maximise the signal-to-noise-plus-interference ratio?'
If any orthonormal beam-forming network is included, as shown in Fig. 13.576,
which transforms the voltages Ex... EN into voltages Ux... UN there is no loss of
information, since all the power incident on the elements is distributed amongst the
outputs. The optimisation problem is then that of choosing the weights oox... coN
and these are clearly related to the original weights Wx... WN by the constants of
the network. The advantage of choosing an eigenbeam network is that the interfer-
ence plus noise components in all channels are decorrelated. The signal steering vec-
tor can be expressed in terms of its eigenbeam components and the problem reverts
to that of combining N channels, each containing coherent signal and incoherent
noise. The solution to this is well known; i.e. the signals should be aligned in phase
and the channels added with amplitude weights proportional to signal amplitude
and inversely proportional to noise power (the 'matched filter' criterion).
At first sight, it may seem curious that the interference plus noise in the eigen-
beam channels can be considered to be decorrelated, since each unique eigenbeam
generally contains power from more than one jammer. The explanation is that it is
simply the long-term average cross-product between the channels which is zero, i.e.

_ I (13.185)

where \t is the fth eigenvalue. For a two-jammer situation, the positive contribution
to this cross-product from one jammer is balanced by an equal negative component
from the other jammer (except, of course, for the self products / = k).
Mathematically, the eigenbeam network shown in Fig. 13.576 is equivalent to an
orthonormal matrix L which is the N by N matrix having the eigenvectors for its
rows.
Thus

L = (13.186)
436 Array signal processing

The relationship between a particular eigenbeam voltage Ut and the element volt-
ages Ei.. .EN is
Ut = J * 1 % + Xi2i)E2 + . . . + X ^ N (13.187)
and hence the vector representing all the eigenbeam voltages is
U = LE (13.188)
Similarly, the steering vector 5* at the elements can be transformed into an equi-
valent steering vector r\ * representing the weights which should be applied to the
eigenbeam channels to obtain the desired quiescent pattern. Thus
TI* = L*S* (13.189)
The same transformation gives the weights co which must be applied to the eigen-
beam voltages U before summation in order that the output of the array shall be
the same as the output of the original array with the element voltages E weighted
by W. Since the output voltage is
ETW = UTv = (LEfu = ETLTo> (13.190)
then T
W = £ co (13.191)
and
co = L*W (13.192)
It will be recalled that the derivation of the maximum signal-to-noise-plus-interfer-
ence condition for the original array:
W = CM~l'S* (13.193)
permitted the elements of the array to have differing patterns. If the eigenbeam
voltages Ux... UN are regarded as the element voltages of a new array, then it will
be evident that the maximum signal-to-noise-plus-interference condition in 'eigen-
beam space' will be of similar form:
co = CM"1-*!* (13.194)
where
U\U2 N 0
¥
U 2Ul U*2UN 0 x2 (13.195)

U*NU2 UNUN 0 0
is the co-variance matrix of the eigenbeam voltages and which, as shown earlier, is
simply the diagonal matrix of the eigenvalues. Since each component of eqn.
13.195 is now decorrelated, the optimum weights in eigenbeam space can be writ-
ten down immediately as

(13.196)
Array signal processing 437

which is the 'matched filter' condition referred to above. The optimum weights W
required for the original array can then be obtained from eqn. 13.191.
To apply the above result to the previously considered case of a 4-element array
illuminated by a single jammer, for which there is one unique eigenvalue:

Qn + 4Pi (13.197)
and three non-unique eigenvalues equal to receiver noise Qni it is first necessary to
choose a set of eigenvectors for the non-unique values. This unique eigenvector is
given by eqn. 13.170, and eqn. 13.173 will be used for the non-unique eigen-
vectors. (It can be shown that the exact choice does not affect the final result and
eqn. 13.174 could also have been used for these eigenvectors.)
The orthonormal transformation matrix representing the eigenbeam former for
this example is then

L = 0.5 (13.198)

which can be recognised as being equivalent to a Butler matrix steered so that one
beam points directly at the jammer.
For simplicity, assume a uniformly weighted steering vector at broadside, i.e.

S* = (13.199)

The transformed steering vector in eigenbeam space is, from eqn. 13.189:

"sin (200

sin + n) oJ(5n/4)
sin(01/2 + ir/4)
n* = o.5 sin (20! + 2TT)

sin (20i + 3TT)

(13.200)
438 Array signal processing
which is simply the set of weights to be applied to the (sin Aft/sin x) shaped eigen-
beams in order to produce maximum gain at broadside. The phase terms in eqn.
13.200 are the result of having chosen the phase centre of the array in the zeroeth
element position and of having selected arbitrary absolute phases for each eigen-
vector. Fig. 13.60 illustrates the weighted eigenbeams for the quiescent condition.
From eqn. 13.196, the optimum weights in eigenbeam space, neglecting common
arbitrary phase and amplitude constants, are:

1 sin (200
\ sin(0i/2)
sin (20i + 7T) eJ (Sir/4)
Qn sin(0!/2 •f it 14)
w = (13.201)
sin (20i •
2}
Qn

Qn
sin(0i/2-
sin (20!
sin (fc/2-
VL )
Since there is only one jammer, the eigenbeam structure is not affected by the jam-
mer power level in this case. Only the <JO\ component of the eigenbeam weight
vector varies if the jammer power changes, since the other eigenbeams have a null in
the jammer direction. Inspection of the equation above indicates that the jammer is
not cancelled completely, but its amplitude is attenuated by the factor
Qn
Qn + &1
The signal-voltage contribution from eigenbeam channel 1 is attenuated by the
same factor. The adapted beam's radiation pattern has a small residual response in
the jammer direction which balances the net signal received from the jammer
against the receiver noise and against the loss of gain in the broadside signal
direction (see Fig. 13.60).
The weights required at the element level to maximise the signal-to-interference-
plus-noise ratio can now be found from the transformation:
W = LTco (13.202)
The result is:

Pi ) (sin(20Q|
W = (13.203)
] (sin(01/2)j

from which it is evident that, even with a single jammer, all element weights have to
Array signal processing 439

jammer

J - = 10 dB

-non-unique Eigenbeams
Eigenbeam
(A, =41)
A* = 1

be adjusted. In the above form, the weight vector is simply the quiescent uniform
weight minus a weighted eigenbeam pointing at the jammer. The (siniVx/sinx) fac-
tor is the amplitude of the quiescent beam in the jammer direction.
Fig. 13.61 shows the result of an eigenbeam analysis for the slightly more com-
plicated case when two jammers are present. In this example, the eigenbeams are no
longer of (siniVx/sinx) shape. The adapted pattern is the same as that shown in Fig.
13.56: the two (sinTVx/sinx) beams in that diagram have merely been resolved into
their eigenbeam components.
440 Array signal processing

quiescent beam

adapted pattern

90°\
JL

F ig. 13.60 Eigenbeam analysis o f adap tive 4-etemen t array with one jammer
a Quiescent pattern
b Eigenbeams corresponding to single jammer
c Eigenbeam components of quiescent pattern
d Adapted pattern and its components

The above examples demonstrate that an eigenbeam analysis can theoretically


reveal exact bearings of jammers (provided / is less than TV), since the non-unique
eigenbeams all have nulls on all the jammer bearings. However, it is somewhat
complicated and in any practical case it requires iterative procedures with an
unpredictable number of iterations in order to determine the eigenvalues. As noted
earlier, it is not used in simple practical adaptive arrays, but a thorough appreciation
of eigenbeams is essential to understanding the basic adaptive processes, particularly
the question of convergence time.
Array signal processing 441

jammer 1 jammer 2
lOr
=
\

/ Eigenbeam 1 . Eigenbeam 2 \
f
\ .*' A 1 = 403.219 A2=38.781

m
V
•10-
•D -1
C
o
en

\l
\

-20

1/ \
-30
180° 90° 90° 180°

(a)

quiescent beam

-A0-
\
• - I weighted . ,
U" I ''-. Eigenbeam 1 II
-50
180°
i ii '"< . ii
90° 0° 90° 180°

(b)

Fig. 13.61 Eigenbeam analysis of adaptive 4-element array with two jammers
a Unique Eigenbeams corresponding to the two jammers
b Adapted pattern and its components
442 Array signal processing

13.73.10 Analogue feedback adaptive arrays:ss An analogue feedback method


of approximately solving the maximum signal to noise plus interference equation
W = CM~l-S* (13.204)
is shown in Fig. 13.62. The signal Vh from element number k is weighted by Wk and
added to all the other weighted signals to form the output I>WkVk. In the absence
of feedback signals, the quiescent weight is essentially determined by the steering
vector component S%. An error signal ek, obtained by correlating the conjugate of
the element voltage V£ with the output, is fed through a low-pass filter which effec-
tively controls the system time constant, and, after further amplification, adjusts
the weight Wk. Wanted signal energy is assumed to be small, and thus Vk consists
largely of interference plus receiver noise.

output 2 (WkVk) = WTV

beam former

Fig. 13.62 Analogue adaptive correlation loop

The output of the A:th loop amplifier is


Wk = G(St-Zh) (13.205)
where Zk is the amplifier input which is related to the correlator output efe, in the
case of a simple low-pass filter, by:
rZk+Zk = ek (13.206)
where r is the filter time constant and the dot means the time derivative. The corre-
lator output for the kth. loop is
N
€k = n z (vkwk) (13.207)

and the outputs of all the correlators can be represented by the column vector
= V*(WTV) (13.208)
Array signal processing 443

where V is the input element voltage vector (noise plus interference) and WT is the
transpose (i.e. corresponding row vector) of the column vector of weights W.
The correlator output vector s can now be written in terms of the co-variance
matrix M as:
€ = MW (13.209)
From eqns. 13.205 and 13.206, ek can be expressed in terms of and its time
derivative Wk by

e _ c._ (13.210)

and hence the complete set of differential equations governing the system can be
written in vector form as
= GS* (13.211)
where / is the identity matrix. Thus the steady-state solution (i.e. when all the com-
ponents of the derivative W have vanished) should approach the desired solution
W = M~X*S* so long as the gain G is sufficiently high.
The manner in which the steady-state solution is reached is not immediately
revealed by eqn. 13.211 because of the interrelationships between the signals in the
various feedback loops. The interdependence can be removed by transforming the
differential equations from the 'weight space' defined by the weight vector W into
'eigenbeam space'. The transformed differential equations can then be solved
readily and the inverse transformation applied, if necessary, to reveal the behaviour
of the actual weights. Paralleling eqns. 13.188 et. seq., let
« = L*W (13.212)
be the transformed weight vector and
(13.213)
be the transformed steering vector, where L is the orthonormal matrix which
defines all the eigenvectors of the co-variance matrix M. Then the transformed
differential equations become
T W + ( G M + / ) oa = (13.214)
where
0 0
X2 0
i = L*MLT = (13.215)
0 X3

0 0 0
has transformed the co-variance matrix into the diagonal form where each element
is an eigenvalue. Each component of the differential equation is now independent:
444 Array signal processing

£>k+{G\k+l)uk = Gr,k (13.216)


and the general solution is

exp ( m i 7 )
(—;—)•
where « 0 = L*W0 and Wo is the initial set of array weights.
The time constant with which the kth component of eqn. 13.217 converges is
r/(GXfe + 1). The value of G has to be chosen such that the loop does not become
unstable for the highest eigenvalue X max . (The greatest loop bandwidth should be
restricted to less than, say, 10% of the signal bandwidth.) The lowest unique eigen-
value X min then determines the rate at which the system finally converges and the
speed at which it can follow dynamic changes in the interference field. Large differ-
ences in jammer power and closely spaced jammers can cause the eigenvalues to be
widely separated and this prolongs the settling time of the system. The larger eigen-
values are cancelled rapidly, but the lower eigenvalues can take perhaps 106 times as
long. Examples of convergence times for single and multiple jammers are shown in
Fig. 13.63.
Convergence of an adaptive array with a single jammer is illustrated by the pat-
terns of Fig. 13.64. The 15-element array is receiving interference from a narrow-
band jammer in the first sidelobe, approximately 11° from broadside. At (a),
shortly after switch-on, the pattern is virtually the quiescent pattern. A little while
later (b) the sidelobe begins to be suppressed and finally a deep null is achieved at
(c).
Figs. 13.65 and 13.66 show the steady state response of a 17-element array to
spatially distributed interference and also show the effect of introducing amplitude
tapering into the array. In both cases the processor has used three zeros to provide
some 50 dB rejection over the region of jamming. In Fig. 13.65 amplitude tapering
has been included in the steering function, whilst in Fig. 13.66 the taper has been
applied at the elements, before the adaptive processor. The sidelobe structure of the
two patterns differs because the steering function in the first case attenuates both
interference and noise from the edge elements, whilst in the second case the taper is
applied to the interference only and the front end noise is unchanged. In the latter
case, less interference power is available to the processor and the eigenvalues of the
covariance matrix are reduced, increasing the convergence time slightly.

13.73.11 Digital implementation of adaptive arrays: The increase in speed of


operation of digital computing equipment over the last few years, together with
rapidly falling costs, have been fortuitous for adaptive array designers. At the time
of writing, the number of digitally-simulated adaptive arrays is several orders greater
than the number of practical digital adaptive arrays in service. The accurate control
of element weights offered by such techniques plus the flexibility of processing
algorithms which can be incorporated are very attractive features which will ensure
the growth of both types!
Array signal processing 445

Gradient-descent methods of implementation are very similar to the analogue


system depicted in Fig. 13.62. One simple technique which is easy to implement is
to perturb each weight in turn (in its in-phase and quadrature components) and
measure the change in output interference power. The gradient vector V can then
be constructed and the weight vector W changed according to the usual iterative
formula:
W W > V f e ) (13.218)

30

DO

15

time —-

Fig. 13.63 Approximate response of adaptive arrays


a Single jammer
b Several Jammers

where k is the iteration number and g is a parameter which controls the stability
and rate of convergence. As in the analogue system, g has to be made small enough
to ensure that the algorithm does not go unstable for the highest expected eigen-
value X max . The disadvantage of this simple step-by-step technique is that it is
inevitably slow. The size of the perturbations has to be restricted to prevent excess
446 Array signal processing

-90 -60 -30 0 30 60 90


ang. from broadside , deg

-60 -30 0 30 60 90
ang. from broadside, deg
(b)

'perturbation noise' and the number of samples averaged for each step has to be suf-
ficiently large to ensure reasonable accuracy in the gradient measurement. Improved
response times can be obtained by modulating all the weights at different frequen-
cies simultaneously and recovering the gradient vector by employing phase sensitive
detectors. Such systems have found application in fixed point-to-point communica-
tion antennas where the interference is essentially stationary.65
Much faster response times are obtained with the widely-used least-mean-square
(LMS) algorithm. This uses digitally implemented versions of correlation detectors
Array signal processing 447

90 -60 -30 0 30
angle from broadside.deg
ic)

Fig. 13.64 15-element adaptive array receiving interference 11° from the broadside (no
amplitude taper)
a Pi/PN = 60dB; G/T = 250; T = 0.1 ms
b PI/PN = 60 dB; G/T = 250; T = 0.75 ms
c Pj/PN = 60dB; G/T = 250; T = 10ms

performing the same function as those shown in Fig. 13.62 to obtain a direct assess-
ment of the gradient vector without needing to perturb the weights. The algorithm
(eqn. 13.218) can be updated at each data sample,66 although some systems average
the gradient measurement over a number of data samples.67 When updating at each
sample is employed, the gain g has to be smaller than 1/Xmax to ensure stability and
low loop noise'. Typically, the effective time constant for the highest eigenvalue
can be as low as 2.57V data samples, where N is the number of array elements. With
a single jammer, the system will then have fully settled in about ION data samples.
However, with multiple jammers the response to lower eigenvalues is more protrac-
ted, the time constant being inversely proportional to the eigenvalue. The perform-
ance with the LMS algorithm is thus similar to that of the analogue array system.
The requirement for algorithms which converge more rapidly for the lower
eigenvalues has stimulated research into efficient methods of performing more
direct matrix inversion.68 In the sample matrix inversion technique (SMI), an esti-
mate is made of the co-variance matrix by averaging successive data samples:

Mk = \ (13.219)
k
and the matrix is inverted by a fast algorithm to compute the weights required.
k = IN samples are usually considered adequate.60'69
In a stationary interference environment, this algorithm can be allowed to run
448 Array signal processing

• spatially distributed
: interference

'-30

-40

-50

-90 -60 -30 0 3


angle from broadside,deg

Fig. 13.65 17-element adaptive array with raised cosine amplitude taper applied to the steer-
ing function
G/T = 500
Pl/PN= 70 dB

: spatially distributed
interference

-90 -60 -30 0 30| 60


angle from broadside,deg

Fig. 13.66 17-element adaptive array with raised cosine amplitude taper applied to the prim-
ary feed
G/T = 500
Pl/PN= 70 dS
Array signal processing 449

from switch-on (k = 1). It proceeds to adapt rapidly, in a somewhat noisy fashion,


becoming smoother as the number of samples averaged is increased. In non-
stationary environments, the algorithm may be updated at intervals of N data sam-
ples, which corresponds roughly to the 'Nyquist' rate required. Alternatively, expo-
nential de-weighting can be employed by updating the estimate according to:
Mk = (l-aW^+aQEffi-i) (0<*<l) (13.220)
where a controls the time constant. Other algorithms can be found69 which update
the inverse matrix M"1 directly at each data sample.

13.7.3.12 Cascade preprocessors: The simpler adaptive antenna systems out-


lined above have a wide spread in convergence rates if the eigenvalues of the co-
variance matrix are widely separated. Although some form of AGC can be applied
to the array as a whole to ensure stability and to ensure that the highest-powered
jammer present is cancelled at the designed convergence rate, the response to lower-
powered jammers is inevitably protracted, the settling time being roughly propor-
tional to the reciprocal of the lowest jammer power. The problem is exacerbated
when the jammer spacings are less than one beamwidth, as this increases the spread
in the eigenvalues.
A method of overcoming this deficiency has been described by White,49 and is
illustrated in Fig. 13.67. The adaptive preprocessor, which can be considered as an
orthonormal beam-forming network of the type shown in Fig. 13.44 (or a digitally
implemented version thereof), resolves the input interference into its eigenvector
components. This is followed by a set of AGC amplifiers which equalise the power
levels. The net effect is that the eigenvalues presented to the succeeding conven-
tional adaptive processor are all equal and the convergence rates for all signal com-
ponents are equalised. The effect on convergence rates is sketched in Fig. 13.67.
The standard processor exhibits a rapid initial cancellation followed by a much
slower cancellation of the lower eigenvalue components. In contrast, the response
with a preprocessor is a classical single exponential decay.
Recursive techniques for controlling the preprocessor are described by White.
Even a single-stage preprocessor which isolates the highest eigenvector alone pro-
vides a marked improvement in performance over a conventional adaptive array.

13.7.3.13 Broadband jamming: The preceding discussion has assumed narrow-


band jamming throughout. The contribution of each jammer to an element mih in
the /th row and fcth column of the co-variance matrix M has been taken as
ife — i k — 1 *^ i u.^<6i^

where Pj is the jammer power and (j>x = (2nd sin Bx)j\.


This representation is adequate only if the variation of phase over the bandwidth
is small, say less than 1°. The phase variation clearly increases as the jammer angle
$i increases. Taking the worst case situation of 6 = 90°, the 1° limit translates into
a fractional bandwidth of
450 Array signal processing

(13.222)
180(M/)
For example, a 10-wavelength aperture array has an 'adaptive null bandwidth' of
0.05% (e.g. 1 MHz a t / = 2000 MHz).

Eigenbeam
r~" former

I'
| I j | AGC amplifiers

i 'standard' adaptive
processor

time

Fig. 13.67 Improvement in convergence rate with cascade preprocessor


Two jammers

To represent broader band jamming, each phase term eHk~l)<i>1 has to be replaced
by a correlation coefficient. Defining the normalised correlation function of a volt-
age E{t) as the complex function
T
lim M E{t)E%t~r)dt
-T
R(r) = (13.223)
1 T
lim — f E(t)E*(t) dt

then the contribution of each jammer to each element of the co-variance matrix
becomes
mik = EfEk = P1R[(i-k)r] (13.224)
where
dsin91
T = (13.225)
A rray signal processing 451

and c is the velocity of EM waves, r is simply the time delay between adjacent ele-
ments for the particular angle Bx. For a single jammer, the covariance matrix is:

Qn PxR(-r) PxR(l -NT)


Pl + Qn PXR{2-Nr)
M = (13.226)
PXR{N-\T) PXR(N~~2T) Qn
The Wiener-Khintchine theorem enables the complex form of correlation function
R(r) to be related to the power spectrum of the jammer G(oo) by the complex
Fourier transform:

R(T) = (13.227)

For example, a jammer with a Gaussian shaped spectrum defined by


1
(13.228)

can be shown to have a complex correlation function

R(T) = exp
-rV exp (JCO0T) (13.229)

and the co-variance matrix can then be written as

Px + Qn
M = (13.230)

Pl+Qn
where
-rV (13.231)
p = exp

is the amplitude of the correlation coefficient between adjacent elements. If this


expression is compared with the co-variance matrix for a narrow-band jammer (for
which p = 1) it can be seen that the amplitude of the coefficients falls off rapidly in
relationship to their distance from the leading diagonal.
The characteristic equation for the eigenvalues
\M-\SI\ = 0 (13.232)
can now be deduced. The phase terms can be eliminated by multiplying every fcth
452 Array signal processing
row by e ®'1^3
and by multiplying every fcth column by e'3^"1^1. The result is:

P
lP n-K PlP
PlP
PlP

(13.233)
Fig. 13.68 illustrates how the eigenvalues vary as a function of correlation coef-
ficient p for a 4-element array with a Gaussian spectrum jammer. For high correla-
tion coefficients (narrow-band jamming) there is one unique eigenvalue and three
non-unique eigenvalues at receiver noise level. As the bandwidth is increased, first
one and then the other low eigenvalues increase. For very broad-band jamming
(p = 0), the eigenvalues become equal.

« or 3"

0.5
correlation coefficient,p

Fig. 13.68 Eigenvalues of a 4-element array with a single broad-band jammer


Receiver noise neglected

It will be clear that in a broad-band jamming environment an adaptive array can


very rapidly use up all its degrees of freedom.
Broad-band nulls can be achieved if time delays, rather than phase shifters, are
used in the array. Fig. 13.69 shows one such configuration,66 which is essentially a
multiloop sidelobe canceller. The main beamformer uses fixed delays to produce a
Array signal processing 453

high-gain beam in the wanted signal direction. Multiple reference inputs to the
canceller are derived from each element. These inputs are fed to a bank of adaptive
transversal filters comprising tapped delay lines with individual control of the
weights on each tap. The combined filter outputs are subtracted from the main
signal. Feedback control of the adaptive transversal filters is obtained from correla-
tion detectors.

main beam-former
(delays)

adaptive transversal
filters

control signals

output
Fig. 13.69 Adaptive array with time delays

13.7.3.14 Summary: The science of adaptive antenna arrays is still developing


rapidly. Much of the research of the past decade has concentrated on the basic
problems of understanding the mechanisms of adaptation and of finding more rapid
and more efficient algorithms for the processing. Latterly, considerable attention
has been paid to the problems of introducing easily implemented constraints to
improve main beam robustness, improving convergence times and improving band-
width of operation.
454 Array signal processing
Whilst considerable ingenuity has been displayed in the mathematics and simula-
tion of adaptive arrays, the bulk of the work has assumed time-stationary noise
fields and the feedback techniques have usually employed simple filters. Future
arrays will have to cope with dynamically varying jamming scenarios, requiring fast-
attack/slow-decay circuits to maintain nulls on the bearings of 'blinking'jammers.
Costs of implementing processors for very large arrays are prohibitive and
current research is investigating techniques such -as sub-arraying which offer the
promise of more economical implementations without too great a sacrifice in per-
formance.
In short, the future for adaptive arrays and for their designers seems assured.

13.8 References

1 FRIIS, H. T., and FELDMAN, C. B. 'Multiple unit steerable antenna', Proc, IRE, 25,
1937,pp.841
2 RYLE, M.: 'A new radio interferometer and its application to the observation of weak
radio sta.is\Proc. Roy. Soc. A., 211, 1952, p. 351
3 VOGLIS, G. M.: 'Design features of advanced scanning sonars based on modulation scan-
ning - Part 1', Ultrasonics July 1971
4 BARTON, P., and KELLY, K. M.: 'Doppler MLS Signal Processing', IEE Conference on
the future of aircraft all weather operation Nov. 1976
5 RADFORD, M. F.: 'The future role of electronic scanning in ATC radar' UK Symp. Elec-
tronics for civil aviation, September 1969
6 RADFORD, M. F., and STEVENS, A. D.: 'A new three-dimensional surveillance radar'.
Military Microwaves 78 Conference proceedings p. 193, Microwave Exhibitions & Pub-
lishers
7 MILNE, K.: The combination of pulse compression with frequency scanning for 3-dimen-
sional radars',/. IERE 28, Aug. 1964
8 MOTKIN, D. L.: 'Practical aspects of a frequency scanning 3D radar'. Conf. Proc. Military
Microwaves, October 1978, p. 203
9 VAN DE LINDT, W. J.: 'Digital techniques for generating synthetic aperture radar
images',/£M/. Res. & Dev. 21,1977, pp. 415-32
10 JACK, M. A., MANES, G. F., GRANT, P. M., ATZENI, C, MASOTTI, L., and COLLINS,
J. H.: 'Real time network analysis based on SAW chirp transform processors', IEEE
Ultrasonics Symposium Proc. 76 CHI 120-5SU, pp. 376-381
11 Proc of conference on 'Optics in radar systems,' Society of Photo-Optical instrumentation
Engineers (Vol. 128) 27-29 Sept. 1977, Huntsville, Alabama
12 SHANKS, H. E.: 'A new technique for electronic scanning',IRE Trans., AP, March 1961,
pp.162-166
13 KUMMER, W. H., VILLENEUVE, A. T., and TERRIO, F. G.: 'Scanning without phase
shifters', Electronics, 29 March 1963, pp. 27-32
14 DAVIES, D. E. N.: 'A fast electronically scanned radar receiving system'/. Brit. IRE, 21,
April 1961, pp. 305-351
15 EDGAR, A. K., and JONES, I. L.: 'Flood-lighting with Nyquist rate scanning', AGARD
1970
16 JOHNSON, M. A.: 'Phased-array beam steering in multiplex sampling', Proc. IEEE, 56,
Nov. 1968
17 WILLIAMS, P. L.: 'Multiple channel receivers for IF beam forming', IEE Conf. Pub. 155.
Radar 77, pp. 427
Array signal processing 455
18 MILNE, K.: Trends in array signal processing*. IEE Conf. Publ. 144. (The impact of new
technologies in signal processing). Sept. 1976., pp. 121-134
19 MILNE, K.: 'Principles and concepts of multistatic surveillance radars'. IEE Conf. Publ.
155 (Radar 77)., Oct. 1977, pp. 45-52
20 CURTIS, T. E.: 'Digital beam forming for sonar systems' IEE Proc, 127, Pt. F, Aug. 1980,
pp.257-265
21 BARTON, P.: 'Digital beam forming for radar', ibid. pp. 266-277
22 MILLS, B. Y., and LITTLE, A. G.: 'A high resolution aerial system of a new type\ Aust.
J. Phys., 6, 1953
23 BLOMMENDAAL, R.: 'A note on multiplicative receiving systems in radar',/. Brit, IRE,
28,1964,pp.317-325
24 SHAW, E., and DA VIES, D. E. N.: Theoretical and experimental studies of the resolution
performance of multiplicative and additive arrays', /. Brit. IRE, 27, Dec. 1965, pp.
323-35
25 DAVIES, D. E. N., and WARD, C. R.: 'Low sidelobe patterns from thinned arrays using
multiplicative processing', Proc. IEE, 237 Pt. F, Feb. 1980, pp. 9-15
26 MOFFET, A. T.: 'Minimum redundancy arrays', IEEE Trans., AP-16, pp. 172-75
27 KSIENSKI, A. A.: 'Multiplicative processing antennas for radar applications,' Radio and
Electron. Eng., 29, 1965, pp. 53-67
28 KSIENSKI, A. A.: 'Signal processing antennas', Microwave J., 4, 1961, No 10. pp. 77-86
(Pt. I) and, No 11. pp. 87-94 (Pt. II)
29 PEDINOFF, M. E., and KSIENSKI, A. A.: 'Multiple target response of data processing
antennas',//^ Trans., AP-10, 1962, pp. 112-126
30 TUCKER, D. G.: The signal/noise performance of electro-acoustic strip arrays', Acustica,
8, 1958,pp. 53-62
31 COOPER, B.C.: The probability density function for the output of a correlator with
bandpass input waveforms',IEEE Trans., IT-11, 1965, pp. 190-194
32 SHEARMAN, E. D. R.: 'Non-colinear and cylindrical multiplicative arrays' /. Brit. IRE,
26, Dec. 1963, pp. 481
33 SHEARMAN, E. D. R., and CLARKE, J.: 'Aperture synthesis in ionospheric radar',
Nature, 219, 1968, pp. 143-144
34 SHEARMAN, E. D. R., BICKENSTAFF, P., and FOTIADES, L.: 'Synthetic aperture sky-
wave radar: techniques and first results', IEE Conf. Publ. (Radar present and future),
London, Oct. 1973, pp. 414
35 HARGER, R. O.: 'Synthetic aperture radar systems', Academic Press, 1970
36 KIRK, Jr. J.C.: 'A discussion of digital processing in synthetic aperture radar', IEEE
Trans., AES-ll,May 1975,p. 326.
37 SKOLNIK, M. I.: 'Radar handbook', MacGraw Hill, New York, 1970
38 BROWN, W. M., and RIORDAN, J. R.: 'Resolution limits with propagation phase errors',
IEEE Trans., AES-6, Sept. 1970, p. 657.
39 PORCELLO, L. J.: IEEE Trans., AES-6, Sept. 1970, p. 636
40 'Remote Sensing of the Terrestrial Environment', Proc. 28th Symposium of Colston
Research Society, Bristol, 5-9 April, 1976, Butterworths, London, 1977
41 CUTTERING, E., et al.: 'Mission design for Seasat-A, an oceanographic satellite' AIAA
77-31, Jan. 1977
42 GUDMANDSEN, P. E.: 'Layer echoes in polar ice sheets', /. Glaciology, 15, 1975,p. 95.
43 PORCELLO, L. J., et.al: 'Apollo lunar sounder radar system', Proc. IEEE, 62, 1974,
p. 769-783.
44 SCHELKUNOFF, S. A.: 'A mathematical theory of linear arrays', Bell System Tech. J.,
22,1943,pp. 80-107
45 DAVIES, D. E. N.: 'Independent angular steering of each zero of the directional pattern
of a linear array,'IEEE Trans., AP-15, Mar. 1967 pp. 296-298
46 HICKS, D. S.: 'Null steering in linear arrays by using amplitude control of signals in the
456 Array signal processing
feeder network/ Electron. Lett., 13, Mar. 1977, pp. 198-199
47 NOLEN, J. C: 'Synthesis of multiple beam networks for arbitrary illuminations.' Bendix
Corporation, Radio Division, Baltimore, Maryland, 21st April 1975
48 WHITE, W. D.: 'Cascade preprocessors for adaptive antennas'. IEEE Trans., AP-24, Sept.
1976,pp.670-684
49 WHITE, W. D.: Adaptive cascade networked for deep nulling*, ibid, AP-26, May 1978, pp.
396-402
50 FIELDING, J. G., et al.\ 'Adaptive interference cancellation in radar systems'. IEE Conf.
Publ. 155 (Radar 77), 1977, pp. 212-217
51 HICKS, D., and RAYMOND, G.: 'Adaptive arrays and sidelobe cancellers for communica-
tions and radar applications'. Conf. Proc. Military Microwaves, MM78, Oct. 1978, pp.
366-378
52 MELLORS, M., DAVIES, D. E. N., and WITHERS, M. J.: 'Zero steering in the directional
pattern of a linear array in the presence of mutual coupling', Proc IEE^ 117, Jan. 1970
53 DAVIES, D. E. N., and RIZK, M. S. A. S.: 'A broadband experimental null steering an-
tenna system for mobile communications', Radio & Electron. Eng., 48, Oct. 1978, pp.
509-517
54 APPLEBAUM, S. P.: 'Adaptive arrays with main beam constraints', IEEE Trans., AP-24,
Sept. 1976, pp. 650-662
55 GABRIEL, W. F.: 'Adaptive arrays - an introduction,' Proc. IEEE, 64, Feb. 1976, pp.
239-271
56 BRENNAN, L. E., PUGH, E. L., and REED, I. S.: 'Control loop noise in adaptive array
antennas,'IEEE Trans., AES-7, Mar. 1971, pp. 254-263
57 GOBERT, J.: 'Adaptive beam weighting', IEE Trans., AP-24, Sept. 1976, pp. 744-748
58 APPLEBAUM, S. P.: 'Adaptive arrays', ibid., AP-24, Sept. 1976, pp. 585-598
59 WIDROW, B.: 'Adaptive filters I: Fundamental', Stanford University Elec. Labs., Syst.
Theory Lab. Centre for Syst. Res. Report SU-SEL-66-126, Technical Report 6764-6,
Dec. 1966
60 HUDSON, J. E.: 'Adaptive array principles', Peter Peregrinus, 1981
61 RALSTON, A., and RABINOWITZ, P.: 'A first course in numerical analysis', McGraw Hill,
1978, 2nd edn.
62 WILKINSON, J. H.: 'The use of iterative methods of finding the latent roots and vectors
of matrices', MTAC, 9, 1955, pp. 184-191
63 WARD, C. R., and HARGREAVE, P. J.: 'The application of sub-optimal control methods
to adaptive antennas for airborne communication systems'. IEE Conf. Pub 195
(Antennas and Propagation) April 1981, pp. 174-178
64 WASYLKIWSKYJ, W., and KAHN, W. K.: 'Scattering properties and mutual coupling of
antennas with prescribed radiation patterns', IEEE Trans., AP-18, Nov. 1970, pp.
741-752
65 WINDRAM, M. D., et al: 'Adaptive antennas for UHF broadcast reception', IEE Proc,
127, Pt. F, Aug. 1980
66 WIDROW, B., and McCOOL, J. M.: 'A comparison of adaptive algorithms based on the
methods of steepest descent and random search', IEEE Trans., AP-24, Sept. 1976, pp.
615-637
67 JENKINS, R.W.: 'Adaptive antenna studies for HF communications'. IEE Conference
Pub. 195. (Antennas and Propagation) Apr. 1981, pp. 190-194
68 HUDSON, J.E.: 'An adaptive antenna with computer assisted convergence'. IEE Conf.
Publ. 169 (Antennas and Propagation) Nov. 1978, pp. 46-50
69 REED, I. S., MALLETT, J. D., and BRENNAN, L. E.: 'Rapid convergence rate in adaptive
mays,9 IEEE Trans., AES-10, Nov. 1974, pp. 853-863
Chapter 14

Radomes
R. H. J. Cary

14.1 Introduction

14.1.1 Definitions
Radomes are defined as electromagnetic windows, consisting of covers or housings
that serve to protect an antenna from damage and environmental conditions, or
modified metallic surfaces such as to permit antennas to radiate. They are required
to have the necessary structural strength, and be so designed as not to exceed some
specified maximum deterioration in the electromagnetic performance of the
antenna under operational conditions.

14.1.2 History and background


The history of radomes (radar domes) largely follows the introduction of micro-
wave antennas, particularly those subject to wind or aerodynamic forces. Since
1940, they have become larger as antennas have increased in size, and more com-
plex in design, due to the demand for higher performance and to the tighter toler-
ances demanded by the use of shorter wavelengths. Various authors have contribu-
ted to the literature describing the evolution, design, manufacture etc., of radomes.
Some of the more important contributions are as follows. Cady,1 describes the
electrical design of normal and streamlined radomes and their installation, together
with the theory of reflection and transmission of electromagnetic waves through
dielectric materials. The accent is on airborne radomes, of solid glaSs^ibre resin
construction, or with cores of foam or honeycomb. Hansen2 concentrates on large
ground radomes, their environmental, structural and design problems, as well as on
methods of construction, including metal spaced frame radomes covered with thin
dielectric materials. Walton3 describes advanced airborne radomes, the problems of
high-speed flight, such as aerodynamic loads, high temperatures and rain erosion,
and discussed materials, e.g. resin fibre products and others suitable for higher tem-
peratures, such as alumina and glass ceramics. Skolnik4 and Jasik5 give theoretical
electrical characteristics for dielectric panels of various constructions and typical
requirements for airborne and ground radomes, with some performance results.
More detailed information is to be found from the Proceedings of the Electromag-
netic Window Symposia at Georgia Tech.,6 and Paris.7
458 Radomes

14.1.3 Constructions
Radomes can be manufactured in a variety of structural forms:
(a) air pressure supported dielectric structure
(b) self supporting dielectric structure
(c) widely spaced metallic or non-metallic frames supporting dielectric windows
(d) (seldon used at present) metal surfaces with dielectric-covered slotted or other
shaped windows.

solid wall

A sandwich
low dielectric constant foam core

A sandwich
low dielectric constant honeycomb core

liiiiiiiiij ihighsandwich
dielectric constant core

C double sandwich

i more than five layers

Fig. 14.1 Radome dielectric constructions

Commonly used constructions, shown in Fig. 14.1, are of the following forms:
(a) A dielectric single-layer wall, either electrically very thin or near multiples of a
half-wavelength thick in the dielectric, such that reflections cancel.
(b) An A-type sandwich consisting of two skins, and electrically very thin or near
multiples of a half-wavelength thick, which are spaced by a lower dielectric con-
stant core of such thickness as to effect substantial cancellation of the skin reflec-
tions.
Radomes 459
(c) A B-type sandwich with outer skins of appropriate dielectric constant and thick-
ness, chosen to match a higher dielectric constant core. This is analogous to optical
blooming, where the outer skins act as quarter-wave transformers.
(d) A C-type sandwich consisting of two A-type sandwiches back to back, forming
five layers, enabling the residual reflections from the individual A sandwiches to be
further cancelled.
(e) More than five layers, with very thin skins and suitable cores, which can give
great strength and wide-band transmission properties for small angles of incidence.
The choice of radome construction will depend on the requirements. A single
'thin' skin may be adequate at low frequencies, but multiples of half-wave thickness
may be necessary to provide sufficient strength at the higher frequencies, where the
sandwich types may be preferred as they give high strength/weight ratio and wide
bandwidths. Metal wires or discs may be included in the construction, e.g. to im-
prove the bandwidth or to match a thinner structure, thereby saving weight.
Ground radomes may be air-pressure supported using inflatable materials, or of
rigid construction, either using solid dielectric, foam dielectric, or sandwich mater-
ials, or using spaced frames with thin dielectric material covers. Airborne and ship
radomes are usually of rigid construction with sandwich or solid dielectric mater-
ials. Underwater radomes are usually of solid materials (Figs. 14.2 and 14.3).

Fig. 14.2 Nose, tail and fin radomes on Hawker Siddley Nimrod aircraft {Courtesy of British
Aerospace)

Dielectric fairings are used to cover suppressed antennas such as slot or notch
antennas energising wings or aircraft tail structures. Radomes are not often em-
ployed with simple unipole or dipole antennas, but rather are used with narrow-
beam microwave antennas, such as scanners on aircraft or large ground-based
antenna systems.
460 Radomes

14.1.4 Disadvantages and advantages of radomes


A radome placed in front of an antenna can interfere with the electromagnetic radi-
ation and will degrade antenna performance, by introducing:

Transmission loss: resulting from the attenuation in the radome wall, reflections
from surfaces, both internal and external diffraction and refraction effects, and
polarisation shift.

Pattern and polarisation distortions: resulting in changes of the main beam-


shape, increases in sidelobes and cross-polarisation of linear or circular polarisation.
Errors in boresight: due to aberrations and phase front distortion of the radi-
ation bv the radome.

Fig. 14.3 Domed cylindrical radomes on HMS Bristol {Courtesy of HM Royal Navy)

The result can be a loss in range performance, reduced bandwidth and tracking
accuracy, an increase in noise temperature and a greater susceptibility to interfer-
ence. This degradation has to be set against possible advantages. With airborne sys-
tems, a radome is often essential to maintain the aerodynamic shape of the vehicle.
With ground and shipborne antennas it has to be established in each case whether or
not a radome is overall an advantage.
The following advantages may be claimed:
Operational advantages
(a) The antenna is not affected by weather, resulting in all-weather operation. How-
ever, the radome itself may be affected by weather.
Radomes 461
(b) The antenna, being enclosed, requires less-frequent and simpler maintenance.
(c) The overall system may be more reliable and accurate.
(d) The working area for personnel is more congenial.
Structural advantages
(a) The antenna structural loads are reduced, due to the removal of environmental
forces. This permits the antenna structure to be simpler and lighter, requiring less
drive power and improved life for the bearings, motors, etc.
(b) The antenna is less prone to corrosion.
(c) The antenna will have more uniform rotation, easier stopping and starting, due
to removal of the environmental forces.
Economic advantages
(a) A cheaper antenna can be produced, requiring less maintenance, but the
additional costs of the radome must be considered.
(b) Small ground antennas, or very large slow-moving or fixed systems, may on
balance favour operating without radomes. But for others, particularly large anten-
nas with high rotational speeds and those operating at the shorter wavelengths, the
balance may be in favour of a radome, provided its electrical degradation, structural
maintenance problems and reliability and cost are acceptable.

14.2 Basic radome requirements

14.2.1 Gen eral requ iremen ts


Radomes have to satisfy aerodynamic, structural and electromagnetic requirements,
as well as environmental and installation problems, such as occur when they are
installed on ship, submarine, ground, aircraft, missile, vehicle etc. These in turn can
call for special material requirements, which can be most severe on high-speed air-
craft and missiles, where aerodynamic heating effects necessitate the use of high-
temperature materials. The electromagnetic requirements could call for certain
limits to transmission loss, aberration, polarisation, pattern distortion, reflection
levels etc., while the structural design must satisfy the shape, size and load require-
ments under all conditions with an adequate safety factor.

14.2.2 Environmental requirements


Typical environmental conditions that may be called for are:
(a) Temperature: Ground radomes — 56°C to + 80°C. Airborne radomes — 56°C to
temperatures dependent on operational conditions and speed, e.g. Mach 2, 135°C,
Mach 3, 280° C. The thermal shock and gradient might also need to be specified on
missiles.
(b) Relative humidity: 100%
(c) Pressure: Ground radomes may have to stand pressures resulting from wind
speeds of up to 240 km/h. Airborne applications demand radome structures capable
of meeting the aerodynamic pressures due to velocity, height and manoeuvre of the
462 Radomes

vehicle, and any internal pressurisation differentials. Underwater radomes have to


withstand the pressure at maximum depth and speed.
(d) Loads: Ice, hail and snow loads have to be allowed for, together with possible
impact loads from birds, stones etc.
(e)Rain: On ground radomes, rain adhesion should be minimised. Airborne
radomes should withstand erosion by rain and hail impact.
(/) Vibration, shock and acceleration: Radomes on aircraft or ships should with-
stand specified vibration, shock and acceleration levels, including explosion and
gunfire if applicable. Ground radomes should withstand seismic loads, e.g. up to 1 g,
depending upon site.
(g) Fire, lightning and radiation hazards: Materials should be fire retardant, and the
radome should be lightning-protected, and able to withstand solar radiation and, if
required, damage from nuclear explosions.
(h) Contaminants: The radome should withstand salt, dust, insects, fungi, fuels,
deicing fluids and other likely contaminants.

14.3 Basic electrical properties of dielectric radome construction

14.3.1 Transmission parameters


The parameters effecting the passage of a ray through a radome are the dielectric
constant, loss tangent and thickness of material, and the polarisation and angle of
incidence of the ray. Cady1 has given formulas to determine the behaviour of the
passage of electromagnetic waves through panels of dielectric material at any angle
of incidence and polarisation, such that amplitudes and phase of the transmitted
and reflected waves, as well as the absorbed energy, can be derived. At normal
incidence the levels are independent of polarisation. At other than normal inci-
dence, multiple reflections and polarisation distortion effects give rise to varying
levels of amplitude and phase of the transmitted and reflected waves, as well as
varying absorption loss.
An alternative approach relies upon the concept of an equivalent transmission
line as shown in Fig. 14.4. A sheet of dielectric then becomes a low-impedance
length of line. A number of layers is a series of such low impedances. Applying
standard matrix representation to each layer (length of line) and performing a
multiplication results in a single representation for the properties of the multilayer.
Thus if the final matrix representing one or a series of layers is of the form
then the various descriptions of the performance of the transparency are:
CD
Voltage transmission coefficient T — -
A+B+C+D

A+B-C-D
Voltage reflection coefficient^ =
A+B+C+D
Radomes 463

COS0
The matrix describing a single layer =
j—sin0 cos0

where the electrical length is 0 =

and d = thickness, X = wavelength, e = dielectric constant and 0 = incidence angle.


Z/Zo is the ratio of impedance in the medium to that in free space, and is:
\fe — sin20 cosfl
and
ecosfl
for parallel and perpendicular polarisation, respectively.
With each layer represented in this form the multilayer matrix is:
A B
= [1] [2] [ 3 ] . . . [«] layers.
B D\
Inductive gratings can be incorporated by introducing a matrix for a shunt
1 0]
susceptance, i.e. I " "I in the appropriate layer. The same formulas can be applied
[JB 1J
to lossy dielectrics by introducing complex values for the dielectric constant and
for the propagation constant.
14.3.2 Transmission of a single-layer construction
In single-layer structures the dielectric constant of materials that have been used
varies from as low as 1.5 on a ground-based foam radome, to as high as 9 for
alumina missile radomes. The most commonly used material is the resin glass-fibre
laminate with a dielectric constant of about 4 and a loss tangent of about 0.015.
The electrical characteristics of such a panel are shown in Figs. 14.5A and 14.5B.
The peaks in the curves are due to the cancellation of the reflections from the inner
and outer surfaces of the panel. The peaks decrease with increasing thickness, due
to the greater loss in the thicker material. The curves are different for perpendicular
and parallel polarisations, and their peak amplitudes and locations are different for
particular angles of incidence. Essentially, this is due to the different impedance
ratios for the two polarisations. Figs. 14.6 and 14.7 show the transmission curves
for alumina with a dielectrical constant of 9 and a loss tangent of 0.0006. The peaks
of transmission occur at consequently thinner physical thicknesses and are narrower
than those of the lower dielectric constant.
The phase delays of the transmissions through single layers of dielectric con-
stants of 4 and 9 are shown in Figs. 14.8 and 14.9. The phase delays vary with
polarisation and angle of incidence, and are greater for higher dielectric constants.
Thicknesses which give maximum transmission or optimum phase character-
istics for a particular polarisation and range of angles of incidence are generally dif-
ferent. A compromise thickness usually has to be selected to achieve a balance
between transmission and least aberration.
464 Radomes

air dielectric air


/

sin 8

equivalent
transmission
lines

longitudinal
2 n cos 8 2n cos 9 propagation
constant

effective
zocos e cos 8 impedance
for Ell

effective
ZQ sec 9 Zosecoc sec 9 impedance
for E l
Fig. 14.4 Transmission through a lossless dielectric
14.3.3 Transmission of an A-sandwich construction
Sandwich constructions comprise a multiplicity of layers of materials of different
dielectric constants. The three-layer A-sandwich with its low dielectric constant core,
has found wide application.1'5 A typical construction is with resin glass-fibre skins
of dielectric constant near 4 and loss tangent 0.015, and core material, such as
honeycomb or foam, with a dielectric constant near 1.15 and loss tangent 0.002.
The transmission characteristics of such a sandwich with skin thicknesses of 0.03
thickness/wavelength ratio, and with variable core thickness is shown in Fig. 14.10,
for perpendicular and for parallel polarisation. The skin thickness chosen is equiva-
lent to lmm. at I band (10 GHz) and forms a lighter structure than the alternative
half-wave thickness of a solid resin glass-fibre construction of dielectric constant 4,
which from Figs. 14.4 and 14.5 would be near 9mm {02S\^x). In consequence,
the weight of the sandwich is about a third of that of the solid single layer. The
Radomes 465

005 010 015 0 20 025 0-30


thickness to free-space wavelength ratio

Fig. 14.5A Perpendicular-polarisation percentage power transmission through dielectric


sheet: Dielectric constant = 4; loss tangent = 0.015

curves indicate reasonable transmission at low angles of incidence with both polar-
isations, but at the higher angles of incidence the transmission is more sharply
tuned, particularly for perpendicular polarisation. Optimum transmission occurs at
widely different core thicknesses for the polarisations. The phase delay through an
A-sandwich is shown in Fig. 14.11. At high angles of incidence there can be large
466 Radornes

differences in delay for perpendicular and parallel polarisation. The A-sandwich is


not usually chosen for operation over a very wide range of angles of incidence, due
to the high levels of reflection losses, and to the large polarisation phase differences
which introduce aberration effects.

100

90

80

c
70
d

Io 60
Q.

cn
d

50

40

30 1 1
005 0-10 0-15 0-20 0-25 0-30
thickness to free-space wavelength ratio
Fig. 14.5B Parallel-polarisation percentage power transmission through dielectric sheet;
Dielectric constant = 4; loss tangent — 0.015.

14.3.4 Transmission of the B-Sandwich construction


An example of a B-sandwich (higher dielectric constant core than skin), is where
the skin is a quarter-wavelength thick and its dielectric constant is the square root
of that of the core, as with optical blooming. This form is equivalent to the quarter-
wave transformer in the transmission-line theory, since each skin and core interface
is matched. The transmission characteristics of such a B-sandwich, whose outer
skins are a quarter-wave thick of resin glass-fibre, dielectric constant 4, with a
titanate-loaded resin glass-fibre core of dielectric constant 16, are shown for perpen-
dicular polarisation in Fig. 14.12. The advantage of this design is that there is more
freedom in the choice of core thickness to satisfy electrical or structural require-
ments, and the harmonic passbands can be useful in multiband application.
Radomes 467
By modifying the core dielectric constant, multiband performance, not neces-
sary in an exact harmonic relationship, can be achieved as shown in the example of
Fig. 14.13, where the dielectric constant of the core is 7 and that of the skins is 4.

100 &

0-025 0 05 010 015

thickness to free-space wavelength ratio

Fig. 14.6 Perpendicular-polarisation percentage power transmission through dielectric sheet:


Dielectric constant = 9; loss tangent = 0.0006.
468 R a domes

14.3.5. Transmission of C-sandwich and further multilayer constructions


The five-layer C-sandwich may be considered as two A-sandwiches back to back. It
can produce maximum transmission at the incident angle where the individual

100

90

80

70

5
1
60

50

30

20 JL _L J_ _L _L
005 0 10 0 15 020
thickness to free-space wavelength ratio

Fig. 14.7 Parallel-polarisation percentage power transmission through dielectric sheet: dielec-
tric constant = 9; loss tangent = 0.0006.

A-sandwiches have zero reflections, and also at the angle where the reflection for
each A-sandwich is of opposite phase. If the individual A-sandwiches give maximum
transmission at one angle of incidence and the combined sandwich at another, then
wide-band coverage over a reasonable range of incidence angles can be obtained. In
consequence, the C-sandwich has found application on high-incidence streamlined
airborne radomes such as that shown in Fig. 14.14. Typical transmission and phase
curves are shown in Figs. 14.15, 14.16 and 14.17 where it may be seen that for a
core of 0.2 thickness/wavelength, high transmission out to high angles of incidence
Radomes 469
is obtained for both polarisations. However, the phase difference of some 40° is
likely to produce significant aberrations.
Sandwich constructions consisting of more than five layers have not been widely
used. Calculations shown that by using thin skin layers and low dielectric constant
cores, useful performance can be obtained, particularly for low angles of incidence,
over very wide bandwidths.

0 01 02
thickness to free-space wavelength ratio

200

80"

] 100 parallel polarisation

0 0 1 0 2 0 3
thickness to f r e e - s p a c e wavelength ratio

Fig. 14.8 Phase delay versus thickness of dielectric sheet at angles of incidence: Dielectric
constant = 4; loss tangent —0.015

14.4 The design of shaped radomes

The input data for any radome design problem is likely to be: the position, size
scan angles and electrical characteristics of the antenna, together with the environ-
mental conditions and loads which the radome will encounter. The desired external
shape of the radome may also be specified, based on aerodynamical considerations,
with the possibility of some minor modification. The initial geometry determines
470 Radomes

perpendicular polarisation

005 010 015


thickness to free-space wavelength ratio

200
80 c

en
<h
parallel polarisation

100
a
.c
Q.

_L _L
005 010 0 15 0 2
thickness to free - space w a v e l e n g t h ratio

Fig. 14.9 Phase delay versus thickness of dielectric sheet at angles of incidence: Dielectric
constant — 9; loss tangent = 0.0006
Radomes 471

0-1 02 0-3 CU 06
core thickness to f r e e - s p a c e wavelength r a t i o

Fig. 14.10 Transmission through A-sandwich with skins of dielectric constant = 4; loss tan-
gent — 0.015 and thickness/wavelength ratio = 0.03; core dielectric constant =
1.15; loss tangent = 0.002.
At various angles of incidence
polarisation perpendicular
polarisation parallel
472 Radomes
200 i

perpendicular polarisation

0-1 0-2 0-3 QL 0-5 0-6

core thickness to free- space wavelength ratio

100-

parallel polarisation

S*
50-

i—
0-1 0-2 0-3 0-4 0-5 0-6
core thickness to f r e e - s p a c e wavelength ratio

Fig. 14.11 Phase delay versus core thickness of A-sandwich with skins of dielectric constant
— 4, loss tangent — 0.015 and thickness wavelength ratio — 0.03; core dielectric
constant = 1.15; loss tangent - 0.002. At various angles of incidence
Radomes 473
the approximate range of angles of incidence. A study of the best choice of mater-
ials, constructional techniques and dimensions that meet the structural and environ-
mental requirements having been made, the resultant electrical performance, trans-
mission loss aberration etc. can be assessed from the transmission equations. By
studying the anticipated performance over a frequency band, dimensions can be
optimised for the particular system bandwidth. It is unlikely that the first optimisa-
tion will be the best solution, and it is usually necessary to vary the radome wall
thickness longitudinally and transversely to refine the electrical design. The best
solution for transmission does not necessarily give the least aberration, and, accord-
ing to the required specification, a compromise in design may be necessary.

20 / \
o
Q_
/ \
c
o

60

Ku
30 GHz
Ka
L35
10 15 20 25

Fig. 14.12 B-sandwich loss X, Ku, Ka bands; perpendicular polarisation


Sandwich construction
Thickness, in 0.045 0.160 0.045
Permittivity 4 16 4
Loss tangent 0.01 0.001 0.01
Angle of incidence

45°
70°

Due to the many variables which affect the antenna-radome combination,


computer-aided optimisation is usually required for the design of a shaped radome.
A typical computer-man interchange flow diagram is shown in Fig. 14.18. Each area
of activity has its own suite of computer programs, for instance, the structural cal-
culations of an airborne radome17 and, similarly, the electrical characteristics of a
radome.21 Furthermore, each suite will contain several subroutines, for example,
cross-polarisation calculation.25
474 Radomes

Fig. 14.13 Modified B-sandwich X, Ku, Ka bands; perpendicular polarisation


Sandwich construction
Thickness, in 0.15 0.07 0.15
Permittivity 4 7 4
Loss tangent 0.01 0.01 0.01
Angle of incidence

45°
70°

Fig. 14.14 Rotating high-incidence C-sandwich AWACS radome (Courtesy of The Boeing
Aerospace Company)
Radomes 475

0-1 0-2 03 <U 05


core thickness to free-space wavelength ratio

Fig. 14.15 Perpendicular-polarisation percentage power transmission of C-sandwich, with


skin dielectric constant — 4, loss tangent — 0.015, thickness/wavelength ratio —
0.03 outer, 0.06 centre, and core dielectric constant — 1.15, loss tangent of
0.002; at various angles of incidence
476 Radomes
14.5 Radome materials

14.5.1 Materials and their requirements


The radome material has to satisfy the structural, electrical and environmental
requirements. Knowledge is required of mechanical properties, such as flexural
moduli, strength, hardness and machinability, together with electrical properties,
such as dielectric constant and loss tangent at various wavelengths and physical
properties, such as specific heat, density, expansion, water absorption, flamm-
ability, and the effects of shock, rain erosion, contamination, radiation, etc.

1001-

90

80

u 70
<u
o
Q.

.60
d

50

J_ JL _L _L
01 02 03 0-4 0-5 06
core thickness to free-space wavelength ratio

Fig. 14.16 Parallel-polarisation percentage power transmission of C-sandwich, with skin di-
electric constant = 4; loss tangent = 0.015, thickness/wavelength ratio = 0.03
outer, 0.06 centre, and core dielectric constant = /. 15, loss tangent = 0.002; at
various angles of incidence

The main materials for air supported radomes are rubberised fabrics, or more
generally Hypalon coated Dacron with a dielectric constant near 2.9 and loss tan-
gent of about 0.008.
Ground, ship and airborne requirements for rigid radomes are largely met by
resin glass-fibre laminates, or sandwiches with core materials of resin fibres honey-
combs or foams. Small radomes may use other dielectric materials such as PTFE,
polyethersulphone etc., with or without fibres. Space frame constructions have
Radomes 477

0-1 02 03 OA 0-5
core thickness to free-space wavelength ratio

150 80c

parallel polarisation
O)
(V

'.100
(V
in
d
CL

50

J_ JL _L
01 02 03 04 05 06
core thickness to free-space wavelength ratio

Fig. 14.17 Phase delay versus core thickness of C sandwich with skin dielectric constant = 4,
loss tangent = 0.015, and thickness to wavelength ratio = 0.03 outer, 0.06 centre,
and core dielectric constant — 1.15, loss tangent, 0.002; at various angles of
incidence
478 Radomes
used resin glass-fibre or metal for the frames, supporting resin glass-fibre laminates
or sandwiches, or skins of organic materials. Missile radomes may call for alumina
or glass ceramics.

14,5.2 Fibre reinforced resin laminates


Properties of common fibre reinforcement materials are given in Table 14.1.

materials electrical
environment construction calculation
input data radome shape thickness of of radome
antenna data radome performance
speed 8> flight
profile if applicable stress
loads
ok

not suitable
change change
material material
thickness thickness

.oads inspected stress loads inspected electrical it ok


and failure for viability and performance out
modes determined weight penalty inspected

not suitable change shape


not suitable change shape

Fig. 14.18 Computer-man interchange radome design

Table 14.1 Properties of radome reinforcements, at 10 GHz and 20° C


Material Dielectric Loss Tensile Young's Specific
constant tangent strength Modulus gravity
S Glass 5.21 0.0078 4 x 10 9 N/m 2 85 x 10 9 N/m 2 2.49
E Glass 6.13 0.0039 3.45 72 2.54
D Glass 4.0 0.0026 2.4 52 2.16
Quartz 3.78 0.0002 1.7 72 2.2
PRD.49III 3.85 0.009 3.45 137 1.45

The fibre yarns can be either knitted, woven or filament wound, with E glass
being the most common, though D glass, quartz and some organic fibres have occas-
sional application where special properties are required. The fibre material is usually
combined with resin by one of the manufacturing methods listed in Table 14.2.
The resultant dielectric constant change with resin loading of a polyester E glass
laminate is shown in Fig. 14.19, and is governed closely by the following empirical
equation:
VR VF
Loge = — — log eF
Y rp Y HP
Radomes 479
where e = resultant dielectric constant
eR = resin dielectric constant
eF = reinforcement dielectric constant
VR = volume of resin
VF = volume of reinforcement
= Vp+ VR+ V, where V is the volume occupied by voids.

Table 14.2 Manufacturing methods for Resin-fibre radomes

Method Advantages Disadvantages


Hand lay-up Offers cheap male and female Labour content high and
tools. Can make complicated demands skill. Accuracy
shapes limited and decreases with
thickness
Vacuum bag As for hand lay-up, but Accuracy demands high skill,
quality better. Slightly more even then is limited
costly
Autoclave Can use pre-impregnated Same as for hand lay-up
cloth, giving accurate repeat-
able products and consistent
glass content
Filament winding Produces high strength, high High dielectric constant
glass content with good leads to thinner walls and
electrical homogeneity. Flex- narrower bandwidth. Wind-
ible for making varying shapes ing machine is expensive.
and sizes Thickness correction may
involve machining outer sur-
face, where, due to low resin
content, the surface is prone
to moisture pick-up unless
sealed
Matched moulds Gives accurate, repeatable Requires high cost outlay
products of high strength, on design, tools and manu-
electrical homogeneity, and facture; hence radomes are
lower dielectric constant than very expensive if only a few
filament wound, giving greater required. Modification of
wall thickness and bandwidth moulding tools to change
radome shape or thickness
may be difficult and
expensive.

The most commonly used resins are polyester and epoxy, with typical dielectric
constants of about 2.7 and loss tangents of 0.013 at 10 GHz, changing with tem-
perature as shown in Fig. 14.20. When combined with glass or quartz, the overall
480 Radomes
dielectric constants are about 4 and 3, with loss tangents of 0.014 and 0.004, res-
pectively. The parameters of a polyester glass composite vary with frequency as
illustrated in Fig. 14.21, and with temperature as in Fig. 14.22<z. A modified poly-
ester8 of lower loss tangent has the property of forming a glass-fibre laminate whose

6i

5-

C
O

o
u 3-

u
<V

2 -

1 -

0 20 40 60 80 100

percentage by weight of polyester resin to E glass fabric

Fig. 14.19 Dielectric constant of E glass laminate versus percentage polyester resin

dielectric constant tends to be independent of temperature (Fig. 14.22&). This is


important for obtaining, over a temperature range, the widest frequency band of
operation. The mechanical properties of typical polyester laminates are shown in
Table 14.3., and of epoxy in Table 14.4.
Radomes 481

29- -0-04

° 28-

002 2

-001

0°C 100°C 200°C


temperature deg.°C
Fig. 14.20 Dielectric constant and loss tangent of conventional polyester resin versus tem-
perature

,£- —
-©-©•
40- •002

•0-015,0

o o
X5^°
3-0- -0-01
•o

25- 0005

0 10 20 30 40
frequency, GHz
Fig. 14.21 Dielectric constant and loss tangent versus frequency (37% polyester SR 17449
resin glass cloth laminate)
482 Radomes

Table 14.3 Typical mechanical properties of polyester laminates

Unmodified Tensile Compressive Flexural Flexural Young's Shear


polyester strength strength strength strength Modulus strength
with MN/m2 MN/m2 MN/m2 GN/m2 GN/m2 MN/m2
E glass 304 266 386 18.6 20 33
pre-preg
E glass 310 280 422 18 20 32
woven cloth
E glass 320 290 490 19
satin fabric
Quartz 200 200 300 16
woven cloth
PRD49III 650 240 300 19 36
woven fabric

0025

002

-0015

c
O

•0005

50°C 100pC 150°C 200°C


temperature, deg.°C

Fig. 14.22 Dielectric constant and loss tangent (10 GHz)


{a) 35% polyester resin and E glass fabric
(/?) 35% modified low loss polyester resin and E glass fabric

The effect of temperatures on the flexural strength of a polyester or epoxy resin


laminate is possibly to lose 10% at 100°C, 25% at 150°C, and 50% at 220°C, de-
pendent on the resin system used (Fig. 14.23).
Radomes 483
Table 14.4 Typical mechanical properties of expoxy laminates

Composite Tensile Compressive Flexural Flexural Young's Shear


and % resin strength strength strength strength Modulus strength
by weight MN/m2 MN/m2 MN/m2 GN/m2 GN/m2 MN/m2
E glass 480 485 630 23
woven cloth
Epon 828
(32%)
E glass 390 390 500 22 24 47
woven cloth
Ciba MY720
PRD49III 608 258 20 36 15
Ciba MY720/
35%
D556 glass 242 364 18 25
Epon 828/
34%
Mil. Spec. 330 344 517 22
5.9300

Higher-temperature performance is obtainable with silicone, phenolic and poly-


imide resins. Silicone resin laminates are of low loss, but have less strength at room
temperature, and are prone to moisture absorption. Phenolic resins are not prefer-
red due to their high loss. Polyimide laminates are favoured for high strength and
high temperature operation, but can be water absorbent unless sealed; otherwise the
dielectric constant and loss tangent can show large increases. The mechanical
properties of a polyimide laminate are shown in Table 14.5. with various reinforce-
ments and the strength degradation with temperature for the E glass laminate is
shown in Fig. 14.24.
These resin fibre laminates satisfy the main environmental requirements, viz,
contamination and solar radiation effects, with temperature limitations, but if sub-
ject to rain impact at high velocity, they can erode. Elastomeric coatings of poly-
urethane or neophrene provide a degree of rain erosion protection. They are usually
applied about 0.4 mm thick, often with a further anti-static paint coating 0.05 mm
thick.

14.5.3 Core materials for resin fibre laminates

14.5.3.1 Honeycombs: Honeycombs are available in a variety of configurations


and sizes using nylon, phenolic, Nomex or polyimide resin-reinforced cells, which
can be formed around doubly-curved surfaces. The electrical properties are shown
in Table 14.6 for nylon phenolic reinforced honeycombs of typical cell size and
density. The loss tangent varies with polarisation, typically between 0.001 and
484 Radomes
Table 14.5 Typical mechanical properties of polyimide laminates
Polyimide Tensile Compressive Flexural Flexural Shear
resin strength, strength, strength, modulus strength,
with MN/m2 MN/m2 MN/m2 GN/m2 MN/m2
E glass 400 360 500 28 15
fabric
Quartz 300 280 300 19
cloth
PRD49IH 350 280 380 25
fabric

600-

500-

100-

o°c 100°C 200°C


temperature, deg. °C

Fig. 14.23 Strength of expoxy resin glass-fibre laminate versus temperature

0.002, at I band (10 GHz), but increases at high incidence angles. Loss tangents for
Nomex resin cells are usually lower.
The mechanical properties of reinforced plastic honeycombs are shown in Table
14.7 for typical call size, densities and resins. Thy polymide has the useful property
of retaining strength to much higher temperatures than most core materials. (Fig.
14.25).
Radomes 485
Table 14.6 Electrical properties of Hexcel nylon reinforced phenolic honeycombs
3/16 in (0.48 cm) cells
Density, Dielectric constant
lb/ft 3 : g/cm2 Perpendicular polarisation Parallel polarisation
E par. L E perp. L E par. L E perp. L
4.5 0.072 1.11 1.07 1.10-1.13 1.07-1.12
6.0 0.096 1.14 1.09 1.13-1.16 1.08-1.14
9.0 0.144 1.20 1.13 1.19-1.24 1.12-1.19

flexural
500-

£00-

compressive

1 300

200-

100

100 200 300 400


temperature^deg C

Fig. 14.24 Strength versus temperature of polyimide laminate (E glass cloth: half hour tem-
perature soak)

14.5.3.2 Foams: A foam core material which has been commonly used is
expanded nitrile ebonite. This has a limited temperature range, softening at 70°C,
and lacks homogeneity. It has been largely superseded by polyurethane, which is
available in various densities and can be accurately controlled to give a close toler-
ance on its dielectric constant. Foams must be of closed-cell construction to avoid
moisture pick up and be homogeneous throughout. This can be achieved when low
486 Radomes

Table 14.7 Typical strength of Hexcell reinforced plastic honeycomb. Samples


1.25 cm thick, cell size 0.48 cm

Type Density, Compressive (stabilised) Plate shear


lb/ft3 Strength, Modulus, (L&W directions)
g/cm3 MN/m2 MN/m2 Strength, Modulus,
MN/m2 MN/m2
Glass 4.0
phenolic 0.062 4.2 390 1.8-1.0 79-35
Glass-nylon 4.5
phenolic 0.07 4.6 550 1.9-0.9 93-36
Glass 4.0
polyimide 0.062 3.0 345 1.9-0.9 200-69
Nomex 4.0
phenolic 0.062 3.9 194 1.7-1.0 63-32

HRH 327: polyimide

30min.
lOOhr.

30min. HRP
phenolic

N
HRP lOOhr. \ 30min.HRH10
* nomex
I 1*0

20-

0 100 200
temperature, deg C

Fig. 14.25 Honeycomb strength versus temperature versus time

molecular weight polyois are used with isocyanate, resulting in a high degree of
cross linkage to give a rigid urethane foam. Water is added during the process to
provide the foaming and to control the ultimate density. Dielectric constant, loss
and mechanical strength all increase with density. In general the material strength
falls off rapidly above 100°C.
Radomes 487
9
A development of a high temperature polyurethane foam, PIO, prepared from
TDI and alkyd triallycyaniviate copolymeric material, has a much higher strength,
it retains its rigidity to 200° C, where in general its strength is still as good as normal
polyurethane foam of the same density at 20°C. The material is closed cell and
softens for forming near 220°C, enabling double curvature shapes to be made.

1-20- -0002

115- - 00015 <o

1-10- - 0 001

105 -00005

1 01
0 005 01 0 15
density, g/cm
Fig. 14.26 Dielectric constant and loss tangent of polyurethane foam versus density

While polyurethane foams are obtainable in almost the whole range of densities,
those typically used for radomes are from 0.05 g/cm3 (3 lb/ft3), to 0.21 g/cm3
(121b/ft3). The dielectric constant and loss tangent variation with density is shown
in Fig. 26.
A comparison of the plastic yield deflection test for normal and P10 polyure-
thane foams of 0.14 g/cm3, (8 lb/ft3), using ASTM Spec. D648-56, is:
Temperature Deflection P10 foam
Normal foam.
20°C 0 0
100°C 0.125 cm (0.05 in) 0
120°C collapse 0
160°C — 0.05 cm (0.02 in)
220°C — collapse
A comparison of strength versus density for normal and P10 foam is shown in
Fig. 14.27. Fig. 14.28 compares their compressive strengths with temperature varia-
tion.
488 Radomes

Polyimide foams are being developed to extend the temperature range to 300°C.

14.5.33. Syntactic foams: An alternative lightweight core material is a syntactic


foam as produced for instance by including low-density glass microballoons10
into resins, offering a choice of dielectric constant less than that of the resin.

301
P10 compressive

20-

10-

normal tensile

2 U 6 8 1b/ft 3
i i
0035 007 0-105 0-U g/cm 3

Fig. 14.27 Strength of polyurethane foams versus density

Typically the dielectric constant of 2.78 and density of 1.2 g/cm3 for an epoxy
resin are reduced to 1.9 and 0.6 g/cm3, respectively, by the addition of micro-
balloons. This low density material, with a flexural strength of 30MN/m2, has appli-
cations in wideband and multiband radome design.

14.5.4 High-temperature materials


For higher temperatures, as encountered on missiles, inorganic materials such as
alumina or glass ceramics11 have been used in solid wall constructions. A summary
of their properties, together with those of other inorganic high-temperature candi-
date materials, is given in Table 14.8. Ceramic foams have been investigated as core
Radomes 489
11
materials for sandwich radomes, particularly alumina and fused silica, as well as
silicon nitride,9 but have presented manufacturing problems in combining the foam
with skins to produce a high tolerance radome.

30 -i

25-
P10

20-

15-

normal

10-

5-

100 200
temperature^

Fig. 14.28 Compressive strength versus temperature of polyurethane foams of 81b/ft3


(0.14g/cmV

14.5.4.1 Alumina (A12O$) The conventional resin glass fabrication has limitations
in operating temperature and in withstanding high-speed rain impact. Alumina is
one of the few current materials which overcomes these failings, and in conse-
quence is widely used for vehicles such as missiles of Mach 3 speed and above.
Electrically, its dielectric constant (about 9.6) is well suited for narrow-band
transmission, particularly where minimum aberration is required. Its loss tangent is
extremely low. Whilst its dielectric constant is nearly constant with frequency, it
changes rapidly with temperature, as shown in Fig. 14.29, limiting its performance
to a narrow temperature range. It can be corrected by adding titanates as shown in
Fig. 14.29, where the resultant dielectric constant is higher but is nearly constant.
The specific gravity of dense alumina, about 3.28 g/cm3, generally results in a
heavy radome, through its high dielectric constant helps to reduce the thickness
Table 14.8 Properties of inorganic radome materials
Property Alumina Pyro- Silica Cordierite Silicon nitride Boron nitride Beryl
i
99% ceram Rayceram Oxide
Slip with Dense Reac- Dense Pyro- 99%
cast Woven tion litic
fused Quartz Bonded
Density g/cm3 3.9 2.6 2.2 1.8 2.45 3.2 2.4 2.0 1.25 2.95
1
lb/ft3 244 162 137 112 153 200 150 125 78 185
X/2 Weight Factor
(g/cm3)/Ve 1.26 1.1 1.2 0.96 1.1 1.15 1.02 0.97 0.72 1.25
20°C10GHz
Dielectric constant
10GHz25°C/43°F 9.6 5.65 3.42 3.05 4.85 7.9 5.6 4.5 3.1 6.6
500°C/832°F 10.3 5.8 3.55 3.04 5.05 8.2 5.7 4.6 3.2 7.2
1000°C/1632°F 11.4 6.1 3.8 3.02 5.8 4.78 3.3 8.0
%e change v.
100°C/180°F 1.2 0.5 0.2 0.1 0.8 0.75 0.4 0.6 0.6 1.5
Loss tangent
10GHz25°C/43°F 0.0001 0.0002 0.0004 0.0009 0.002 0.004 0.001 0.0003 0.0003 0.0003
500°C/832°F 0.0005 0.001 0.001 0.001 0.008 0.0045 0.0025 0.0006 '0.0006 0.0005
1000°C/1632°F 0.0014 0.0008 0.0014
Flexural strength
25°CN/m 2 Xl0 6 270 235 44 125 400 100 100 260
500°C 250 200 54 120 400 60 200
1000°C 220 75 66 100
43°Flb/in 2 Xl0 3 40 34 6.3 18 57 25 14 14 35
832°F 37 29 7.8 17 57 9 29
1632°F 35 11 9.5 14
Young's modulus
25°CN/m 2 X 109 380 120 48 18 128 300 70 13 320
500° C 350 120 48 125 300 50 300
1000°C 285 100 120 210
Table 14.8 Continued
Property Alumina Pyro- Silica Cordierite Silicon nitride Boron nitride Beryl
99% ceram Rayceram Oxide
Slip with Dense Reac- Dense Pyro- 99%
cast Woven tion litic
fused Quartz Bonded
Young's modulus
43°Flb/in2X 106 54 17 7 2.5 18 43 15 10 1.75 46
832°F 50 17 7 — 17.6 43 7 42
1632°F 41 14 17 30
Poisson ratio 0.07-
0-800°C 0.28 0.245 0.15 0.18 0.26 0.23 0.34
Coef. thermal
conductivity
BTU.ft.h°F 20 2.2 0.46 2.4 1.4 12 6 14 16 120
cal.cm.s°C 0.09 0.009 0.0019 0.01 0.006 0.05 0.06 0.07 0.5
Coeff. thermal
expansion
10-6in/in/°F 4.5 2.5 0.30 1.3 1.8 1.4 1.8 2.1 4.6
10-6cm/cm/°C 8.1 4.0 0.54 2.3 3.2 3.2 3.8 8.3
Specific heat
cal.gr.°C 0.27 0.2 0.18 0.25 0.18 0.2 0.2 0.3 0.29 0.26
Thermal shock Fair Good Very Very Good Very Very Very Very Very
good good good good good good good
Water absorption 0% 0% 5% 26% 0% - 23% 0% Closed 0% 5
cell
Rain erosion Excel- Very Poor Poor to Very Very Good Very Good Very
!
lent good fair good good good good
492 Radomes

required. Alumina is one of the hardest materials, and in consequence is extremely


good for resisting rain erosion, but on the other hand it is difficult and costly to
grind to shape. Its thermal expansion coefficient is high, and the radome is thus
prone to thermal shock. As a guide, 300°C thermal shock is typically considered
safe, but a 400°C differential may result in failure. However, much will depend
upon the missile's flight envelope and environmental conditions, and the radome
shape and thickness. The mechanical properties of dense alumina with temperature
variations are shown in Fig. 14.30. Alumina can be manufactured to a high degree
of purity and density, thereby ensuring homogeneity and consistent electrical
characteristics. Densities of 95 to 99.5% are obtained, according to manufacturing
method, the most successful using pressing processes. The variation of dielectric
constant with density is shown in Fig. 14.31.

titanate alumina

11-

99H-7o

10-

500 1000

temperature,deg C

Fig. 14.29 Dielectric constant of alumina and titanate alumina versus temperature

14.5.4.2 Glass ceramics Glass ceramics are the other main candidate inorganic
materials for radomes requiring high-temperature capability. The glass ceramic
mainly used is Pyroceram 9606, u a product of the Corning Glass Co., though other
variants are available with various emphases on particular properties.
Radomes 493

1500-

compressive

1000-

300-

flexural

en
S 200

tensile

100-

500 1000
temperature ,deg C

Fig. 14.30 Strength versus temperature of dense alumina

9 5-

90-

8 5-

3 6 3-7 38 39
d
density, g/cm

Fig. 1431 Alumina dielectric constant versus density at room temperature and 9.5 GHz
494 Radomes

Pyroceram has a dielectric constant of 5.65, changing little with frequency or


temperature variations. (Fig. 14.32). This enables designs to be achieved with high
transmission and reasonably acceptable aberration characteristics over a somewhat
wider band than alumina radomes. The loss tangent is low at ambient temperature,
rising to a still acceptable value of 0.01 at 900°C.

.a 5 5-

10 20 30
frequency, GHz

55-

500 1000
temperature.,deg C

Fig. 14.32 Dielectric constant versus frequency at 20° C of Pyroceram 9606 and versus tem-
perature at 9.5 GHz

Mechanically the material is hard, with good rain erosion resistance, and thermal
shock properties. Its specific gravity of 2.62 g/cm3 makes it lighter than alumina,
but for the half-wave thicknesses radome design the net weight is only slightly less
than that of alumina. Some of its mechanical strength properties, Fig. 14.33, are
lower than those of alumina, but on balance, due to the increased physical thick-
ness of the half-wave sheet usually chosen for the design, its properties are
adequate.
A glass ceramic material based on Zn, Al, Si and Ba oxides, called Mexim, has
been fabricated into radomes by GEC (UK) with a dielectric constant varying
according to formulation, but typically near 6 with a loss tangent of 0.001 at
500°C. Mechanically and thermally, the material has properties somewhat similar to
Pyroceram.

14.5.4.3 Silica and reinforced silica Silica is another high-temperature candidate


Radomes 495
material, though little used to date. Silica can occur as quartz, and in the less
dense form of fused silica, as developed particularly at Georgia Tech. (SA).11 Its
lower dielectric constant, around 3.5 depending on density, (Fig. 14.34) has the
advantage of giving low reflection losses over much wider bands than alumina
radomes. Its dielectric constant changes little, over a wide temperature range. (Fig.
35). The material has a low loss tangent, about 0.004 at 10 GHz. Mechanically,
silica is weaker than alumina or Pyroceram. (Fig. 14.36). Silica can produce clear
quartz, when heated to its liquid stage, and as such is used for small radomes. Fused
silica, formed by a sintering slip casting process, is the more usual process of manu-
facture and, due to its porous nature, methods of preventing water absorption have
to be incorporated. The material has a poor rain erosion resistance.

1500 -
compressive

1000 -

300
CM

E flexural

2
200
_c
"en
c
<b

100

0 500 1000

temperature, deg °c

Fig. 14.33 Strength versus temperature of Pyroceram 9606

14.5.4.4 Other high temperature materials Combinations of oxides, such as


cordierite, mullite, streatite, spinel, beryllium and glass aluminium phosphate, have
some potential as high-temperature materials, but do not have better overall charac-
teristics than alumina or glass ceramics.
The nitrides of silicon and boron are considered as possible contenders for
radomes which require high thermal shock capability. Silicon nitride in particular
has suitable electrical and mechanical properties. The dielectric constant of dense
silicon nitride (3.2 g/cm3) is near 7.9, and for the less dense reaction-bonded variety
496 Radomes

u, 3 75 -

325

2 21
density , g/cm 3

Fig. 14.34 Dielectric constant versus density of silica at 9.5 GHz and 20° C

dense

slip cast

3 5-

0 500 1000
temperatureJdeg°C

Fig. 14.35 Dielectric constant versus temperature of silica at 9.5 GHz


Radomes 497
200 H

very dense
t 100H

less dense

0 500 1000
temprature,deg C

Fig. 1 4 3 6 Flexural strength versus temperature of silica

max density

25 2-75 30 325
density, g /cms

Fig. 14.37 Dielectric constant versus density of silicon nitride at 9.5 GHz and 20° C
498 Radomes

(2.5g/cm 3 ) it is near 5.5 (Fig. 14.37). It changes considerably less than that of
alumina with temperature (see Fig. 14.38) and the material exhibits low loss. The
mechanical strength changes little with temperature (Fig. 14.39). Rain erosion resis-
tance is good, and improves with increasing density. Thermally, the shock resistance
is excellent, which is the main advantage. The densest material requires heavy
pressures and costly tools for its manufacture, while the reaction bonded silicon
nitride is porous, presenting sealing problems.

9 -

_L _L
500 1000
temperature , deg °c

Fig. 14.38 Dielectric constant versus temperature of silicon nitride at 9.5 GHz

14.5.5 Loaded and artificial dielectrics


These have application where the permittivity of a material is required to be
increased to a particular value. This can be achieved by adding a high-permittivity
material, e.g. a titanate of e = 90, in suitable quantities. Alternatively, the inclusion
of metallic particles of various shapes, small compared to the wavelength, has been
shown to produce a higher permittivity artificially without noticeably increasing
the weight. Moorefield12 has described aluminium-covered hollow glass spheres
added to a foam to produce a 3.2 dielectric constant lightweight core material
matching resin quartz fibre skins.
The effective dielectric constant of homogeneous distributions of polarisable
conductive elements in air can be shown to approximate to the formulae given in
Fig. 14.40 for metallic spheres, discs and rods. Irregular arrangements of metallic
elements result in lower dielectric constants due to the lower polarisability arising
from the random orientation of the elements. Measured dielectric constants at
10 GHz of various forms of artificial dielectrics in foam of dielectric constant 1.2
are:13
Radomes 499
Discs: diameter lmm, spaced 8 mm, layer spacing 0.8 mm, dielectric constant =
2.6, loss tangent 0.11
Dipoles: length 2.5mm, spaced 0.3mm, layer spacing lmm, dielectric constant =
3.2, loss tangent 0.015
Aluminium flake with increasing random loading
dielectric constant 1.5 2.0 2.8
loss tangent 0.01 0.015 0.03

500 -

dense

300

5 200
reaction bonded

100

J_
0 500 1000
temperature, deg °c

Fig. 14.39 Flexural strength versus temperature of silicon nitride

There can be problems in the manufacture of artificial dielectric materials, as


measurements often show a considerable spread in dielectric constant and loss
tangent, particularly at the higher loading densities.

14.5.6 Rain-erosion resistant materials


Airborne radomes have a special problem of rain erosion.14 Some comparative
results on erosion of materials on a whirling arm rig, at 500mile/h in 1 in/h rainfall
are indicated in Table 14.9. The figures need interpretation for practical airborne or
missile radomes, depending on the speed of the vehicle, the angle of incidence,
flight envelopes etc. Clearly alumina and Pyroceram are in a class of their own,
compared to glass-fibre laminates.
Application of a polyurethane or neoprene coating, recommended to be at least
500 Radomes
Table 14.9 Comparative rain erosion rates
Material Effect after Time
Alumina (dense) no erosion 4200 min
Pyroceram (9606) slight pitting 2586 min
Silicon nitride (RB) deep pitting 105 min

Glass resin laminates


polyurethane 0.4 mm coat damage beginning 80 to 120 min
neophrene 0.4 mm coat damage beginning 25 to 40 min
uncoated damage beginning 5 to 15 min

a = sphere radius

s = sphere spacing

= 14-16a°
3S 2 L
= disc radius

disc spacing

layer spacing

1-1

volume fraction of
elements in medium

/ length \
\ diameter /
n /f 2 length \

Fig. 14.40 Permittivity of regular spheres, discs and rods of metal


Radomes 501
0.4 mm thick, is obviously advantageous to the life of glass-fibre laminates, and the
thicker the coatings the longer is the time of survival. These coatings are tempera-
ture-limited, with polyurethane being capable of working to a higher temperature
(at least 150°C). It is usually recommended both for this reason and because of its
longer life. The dielectric constant of the coatings of about 3.2 is a fair match to
the usual fibre resin products to which they are applied, but their high loss
tangents, near 0.05 at 20°C and increasing with temperature, become significant
limitations at the shorter wavelengths.

14.6 Radome structures

Structural design can be a very complicated procedure and may require many com-
promises to be made before all aerodynamic, electrical, operational and environ-
mental requirements are met. The basic approach, after selecting a preliminary
design, is to determine the loads and evaluate the structural integrity and, if unsatis-
factory, modify the design until an acceptable solution is found. It is a fundamental
requirement that the structure be safe; hence the flow diagram of Fig. 14.18
attempts to find such a safe structure which is closest to meeting the electrical
requirements. Simplified, the structural design may consist of first evaluating the
loads, taking into account the appropriate wind, aerodynamic and blast pressures,
inertia, shock, temperature range stresses, ice, snow etc.; secondly examining the
materials and configurations that might meet the mechanical and electrical require-
ments; and thirdly calculating the dimensions of the materials, and possibly modify-
ing the radome shape to best advantage to satisfy the structural, aerodynamic and
electric requirements.
The non-rigid radome, relying for support on internal air pressure, requires simi-
lar consideration. These radomes are usually ground based and the important loads
which enable the internal air pressure to be determined are those due to wind, ice
and material weight.

14.6.1 Near-spherical rigid shapes (usually on land or sea)


If the radome shape is rigid, the Tsien formula gives the uniform buckling pressure
Pas:

where k is normally taken as 0.3


E = elastic modulus of the material
R = radius
t = thickness
In an air flow, the molecules directly ahead of the radome tend to part and
streamline around the surface. The total energy is the sum of kinetic, potential and
pressure energies possessed by the air, and at one point is the same energy as at any
502 Radomes

Fig. 14.41 Pressure diagram of air flow round sphere


Radomes 503

other point. The air flow around the sides of the sphere is constricted, with the
streamlines closer together, causing a reduction in pressure. The result is that the
pressure pattern over the body may take the form showin in Fig. 14.41, with both
positive and negative pressures. If the maximum pressure, e.g. at the stagnation
point, is one half the air density times the velocity squared, then the required thick-
ness can be solved from buckling theory.
With a space-frame radome,15 the assembly of beams and panels effectively acts
as a shell, whose effective modulus and thickness is given by:
EA
Ee
_ and te
ft^~
" 467 jnT ~~ v T
\Jn N A
which by substitution into the buckling equation gives:
/nAI
P = 0.225£-
R3
when E — elastic modulus of framework beam
/ = maximum moment of inertia of beam
A = beam cross-sectional area
R = radius
n = number of panels in spherical radome
Hence if the pressure is known the minimum beam cross-section is determined. In
addition to evaluating the general stability of the radome framework, it is necessary
that the individual beams do not buckle under the design loads, and that the
strengths of both the membranes and the beam to membrane attachments are ade-
quate. The whole is best achieved by computer analysis.

14.6.2 Non-spherical rigid shapes (airborne applications)


Airborne radomes are shaped to avoid drag, and are often far from spherical,
notably nose radomes. In Fig. 14.42, a symmetrical pressure pattern is shown on a
radome nose with positive stagnation pressure head on, decreasing to negative pres-
sure aft. A section of the radome gives rise to a hoop stress of pR/t, where p is the
pressure on the section, R is the radius, and t is the thickness. Unsymmetrical pres-
sures are produced by the vehicle manoeuvre, such as yaw, banking, impact, vibra-
tion etc. This results in longitudinal and circumferential compressive, tensile and
shear stresses varying along the radome and with speed, altitude etc.
A typical pressure load pattern for one condition of flight is shown in Fig.
14.43#, from which stresses have to be determined. A family of such patterns under
all loads has to be analysed for all flight conditions. The integrated local loading
round the periphery per unit longitudinal length at radius r at distance x back from
the nose is: L(kN/m) = f^prdO; e.g. in Fig. 14.435, 1 m back from the nose, there
is an upload of 17 kN/m. Similarly the maximum up, down and side loads are deter-
mined from the various pressure distribution diagrams at maximum velocity and
manoeuvre. The pressure loading curves permit shear and bending moment diagrams
504 Radorn es

to be obtained. The shear force diagram is found from S = f£ Ldx, e.g. at the lm
station back from the nose, the shear load, shown in Fig. 14.43&, is near 8 kN. The
bending moment diagram is similarly determined from BM = f$Sdx, and at l m
station, the bending moment, as shown in Fig. 14.43&, is near 2kNm. These dia-
grams, together with other forces, enable the total stresses to be calculated and the
failure modes determined.

—S. = hoop stress

longitudinal section uniform pressure


compressive and tension stress on cross section

non-uniform pressure
on cross section

Fig. 14.42 Symmetrical and asymmetrical pressure diagrams on nose radome

The choice of shape, material and construction of the radome available to the
structural designer may be limited by microwave or aerodynamic requirements. For
a nose radome, a procedure may be to consider the radome as a shell, comprising a
series of rings which each have to be strong enough to carry their individual loads.
Radomes 505
r- 2m. up surface

2m along surface
J from nose apex

stagnation pressure
115kN

- 2 m down surface from nose apex

local aerodynamic
pressure loading, kN
20- m

shear force, kN

bending moment., kNm

10-

T I
0 1 2
distance from nose apex, m
Fig. 14.43A Aerodynamic pressure distribution on nose radome (maximum upload condition)

Fig. 14.45B Pressure distribution and loads on nose radome (maximum upload condition)
506 Radomes

The inertial loads, when integrated, are likely to be greatest at the base of the struc-
ture, and lead to an analysis of the base attachment, for shear and bending
strengths.
To save time, load points, which are considered likely to cause failure are inspec-
ted first. Primary shell failure modes are tensile fracture and compression insta-
bility. Tensile fractures may be simple, a creep rupture, or a fatigue crack. Simple
rupture occurs when applied loads cause excessive tensile stress. Creep rupture is
time dependent and occurs when a material continues to stretch under load until
failure occurs; this tends to happen with resin-bonded materials. Fatigue crack
occurs when loads are applied a number of times, and is more evident with brittle
ceramic materials. Instability failure occurs when the structure collapses under direct
compressive stress, as can happen with thin-walled structures. Local instability can
occur with resin glass-fibre laminates, due to delamination, bond or cure failure,
and with inorganic compounds, such as alumina, due to local gradient stresses.
With advanced radomes the analysis of shell structures becomes very complex
when using realistic loading patterns and constructional details. To cope with this
the radome is often analysed with aid of computer programs, generally using finite
elements17 (Fig. 14.44). Structural elements commonly used are rods, beams, rings,
membranes, conical and cylindrical shells and flat plates. Mixtures of these simple
elements may be used to approximate to structural details. A coarse network of
elements may be used in a preliminary analysis to minimise computation; this will
often suggest changes to the model to improve the stress distribution. The input
data required is:
(a) geometric data giving co-ordinates of connecting points of the structure ele-
ments
(b) boundary conditions of location and type of constraints acting on the network
and the type of load interactions which may occur between elements
(c) stiffness data for the type of element, its cross-sectional dimensions, variations
of properties through the cross-section, and appropriate integrated stiffness para-
meters of cross-section, (moment of inertia of beam, area of rod, etc.)
(d) loading data such as pressure, inertia, temperature variations over and through
the cross-section, concentrated loads, etc.
In spite of the confidence engendered by comprehensive computer-aided predic-
tions of actual stresses, the design is usually verified by practical tests on the
radome under simulated loading and environmental conditions, to confirm the
radome's safety factor. Some commonly performed structure tests are:
(a) instability tests under static and impulsive loads
(b) dynamic tests under harmonic and random vibration and loads.
If the failure mechanism is not easily quantifiable, for example the effects of
rain or hail impact and erosion, high-temperature charring of the surface, thermal
shock resistance, etc., then functional tests should be performed on specimens to
establish likely failure effects.
Radomes 507
14.7 Airborne radomes

Airborne radomes may vary widely in size and shape, may call for near normal
and/or high incidence-angle transmission, and may be used on low- or high-speed
vehicles in a variety of flight and environmental conditions. Transmission data of
the form referred to in Section 14.3., together with structural prediction tech-
niques, are used to find solutions to the requirements. Radome requirements may
include not only adequate transmission, but wide bandwidths, minimum aberration,
low crosspolarisation and low reflection.

Fig. 14.44 Aircraft nose radome finite element model (Reference 11}

14.7.1. Wide-band transmission designs

14.7.1.1 Wide-band single-layer and sandwich designs The single-layer radome


has maximum transmission when the thickness is very small compared to the wave-
length or when the reflections cancel at the half-wavelength (in the dielectric) thick-
ness, and at multiples thereof. Further, the lower the dielectric constant, the lower
is the reflection; and the lower the absorption loss, the greater is the transmission.
A single layer of a thin skin of 0.5 mm thickness might be structurally practical,
for example, if the criterion is for angles of incidence up to 60°. From Figs. 14.4-
14.7, the bandwidths covered would be 0 to 3.2 GHz and 15 to 17 GHz for a di-
electric constant of 4, and 0 to 1.5 GHz, and 7.5 to 8.0 GHz for e = 9. (The upper
passbands are where the radome becomes a half-wave thick.) At the shorter wave-
lengths, structural reasons may demand a half-wave thick design, or even a full-wave
thickness, which results in a further reduction in bandwidth. The bandwidths are
mainly limited by the performance in perpendicular polarisation. If the requirement
was of 0° to 80° angle of incidence coverage, the performance would be further
reduced.
In practice, the bandwidth achievable may be further reduced by changes of
dielectric constant over the temperature range and by variations of loss tangent. The
508 Radorn es

bandwidth improvement over a temperature range of — 40°C to 150°C obtained by


using the modified low loss, temperature-stable dielectric constant polyester (Fig.
14.22) in a laminate, instead of the conventional type is indicated in Fig. 14.45, for
80% transmission over angles of incidence up to 80°. Some further improvement
can be obtained by using a lower dielectric constant quartz cloth (b) instead of the
E-glass reinforcement (a).

.with low loss


temperature stable
polyester

50° 60° 70°


angle of incidence, deg

Fig. 14.45 Bandwidth improvement of resin laminates with low-loss temperature-stable poly-
ester
(a) with E glass
(b) quartz reinforcement

The sandwich type of radome relies on maximum transmission when each skin is
spaced such that reflections cancel. The bandwidth of the A sandwich can be deter-
mined from a presentation such as that shown in Fig. 14.46#, where constant
power transmission curves for perpendicular polarisation for particular angles of
incidence are shown for varying core and skin thicknesses. Parallel polarisation is
not shown as it usually gives higher transmission up to the Brewster angle.
Operation over angles of incidence 0° to 60°, with greater than 80% transmission, is
obtained along the line 0 to A, (Fig. 14.46a), with a core up to 0.4 wavelengths
thick and corresponding skins up to 0.03 wavelengths thick. If the band was to
cover all wavelengths down to 3 cm (0-10 GHz), the sandwich would have a core of
1.2 cm and skins of 0.9 mm thickness.
Radomes 509

\^50° 70°

S0.3 • \ 70°^^*^
vei

\ ^v. ^^^-

o 70° \ ^^-^
**—^^-22°
— ^ •^^__o°
80.2-
c \ / —
" " ^ "^ —
D
__ 50°
i o.i •
o
rati

/ 3Q2- -
/ \ /
0
0.01 0.02 0.03
skin thickness
wavelength

0.4 I

0.03
skin thickness
wavelength

Fig. 14.46 Angles of incidence curves of 100 and 80% transmission of perpendicular polaris-
ation of (a) lossless A and (b) C sandwiches (a: skin, 4; core, 1.15)
100%
80%
510 Radomes
If a reduced bandwidth of 4-10 GHz was required, but with angles of incidence
up to 70°, the bandwidth would be determined from C to B, (Fig. 14.46a), with a
thickness to minimum wavelength ratio of 0.3 for the core and 0.025 for the skin.
If very high angles of incidence were required, it would be necessary to include a
study of parallel polarisation, which, beyond the Brewster angle, starts to limit the
performance of wideband A sandwiches.
The C sandwich has the advantage that it can have perfect transmission at more
than one angle of incidence, e.g. point P in Fig. \4A6b, where the 100% transmis-
sion curves for 0° and 60° angles of incidence intersect. The curves indicate that
wide-band transmission properties can be obtained, even at high angles of incidence.

14.7.1.2 Inclusion of metallic elements to increase bandwidth Metallic elements


can be included to increase the bandwidth. Since the effective susceptance of a
dielectric sheet less than quarter-wave thick is capacitative, it can be compensated
by a grid of wires which are inductive. The normalised susceptance of an air/
dielectric interface, viewed from within the dielectric, for perpendicular polarisa-
tion incidents, is:18

where tan 0 = tan

e
~

I = angle of incidence
e = dielectric constant
d = distance back in dielectric from interface
X = air wavelength
whilst for parallel polarisation
__ eW/
e — sin2/
The susceptance of a grid of parallel wires embedded in a dielectric sheet where
the electric vector is parallel to the wires, and the polarisation is perpendicular to
the plane of incidence, is:

B —
+ 0.6 I - 1 (e + 2sin2/)
\ A/

where P = pitch of wires


D = diameter of wires
/ = angle of incidence
e = dielectric constant of sheet material
Radomes 511
2 2
For parallel polarisation the term, e 4- 2sin /, is replaced by, e — sin /. In Figs.
14.47 and 14.48, susceptances of various diameters and pitches of grids of wires are
shown for 0° and 80° angles of incidence in dielectric constants of 1.15 and 4,
which are typical of resin glass solid and sandwich constructions. The susceptance
of grids in high dielectric constant materials is almost independent of the angle of
incidence.
A wide bandwidth may be obtained, e.g. by arranging the susceptances of the di-
electric sheet and those of the wire grid to cancel at two frequencies, as shown in
Fig. 49, where a 4:1 frequency band is obtained with low transmission loss.
An example of matching a high angle-of-incidence ogive-shaped radome, while
giving little ellipticity for a circular polarised transmission, was given by Cary.19 The
inductive matching took the form of circumferential coils spaced along the radome,
set in the centre of an A sandwich of resin quartz-fibre skins, of dielectric constant
3.2, and a foam core 1.2. The radome could be matched theoretically either by a
practical size, 0.05 mm diameter, of circumferential spaced wires, but at such a
spacing to give rise to diffraction lobes, or by an impracticable 0.002 mm diameter,
wire grid closely spaced to avoid diffraction lobes. This problem was solved by
introducing some capacitance to the wires by coiling a 0.05 mm-diameter wire at
one turn per mm on to a thin-walled 0.84 mm-diameter tube, such as to give the
required net inductance, equivalent to that of the impractical wire size. The range
of angles of incidence was mostly from 0° to 72° as the antenna scanned. From cal-
culation, it was expected that the coiled matching of the A sandwich for circular
polarisation would not give more than 18% transmission loss at a low frequency in
the microwave band, and would rise to 23% at four times the low frequency, and
that the ellipticity ratio would not be worse than 0.7. The practical result con-
firmed this performance over ± 30° antenna scan (Fig. 14.50).

14.7.2 Multi-band radome designs


The single-layer radome gives passbands when it is very thin compared to the wave-
length, and at half-wave thickness and at multiples thereof. Inclusion of metallic
elements can give multi-band operation, by arranging the elements to be virtually
invisible at the normal radome operational band, but of such a susceptance as to
match the radome at another band. A radome required to be matched at both I
band (10 GHz) and D band (1 GHz) can be designed by choosing the radome to be a
half-wave thick at I band, but to include a grid of wires electrically widely spaced
such as to be of negligible susceptance at this frequency. At D band, the radome
will appear electrically thin and the grid to be closely spaced, such that the two
susceptances cancel and the radome is again matched.
The A sandwich can be adapted to give multi-band operation. For instance, the
skins can be chosen to be a half-wave thick at a short wavelength, but at a longer
wavelength the skins would be electrically thin, and if spaced a quarter wave apart
would give another pass band. At even longer wavelengths there would be a low-
frequency passband, as the whole sandwich would be electrically very thin. An
advanced airborne nose sandwich radome has been described,20 with half-wave skins
of 0.86 cm, so spaced as to give passbands at 0 to 1, 10 and 36 GHz.
512 Radomes

S 10
"09
08
0 7
06
05

0U

03

0 2

0 1
01 0 2 0 3 04 05 06 07
wire pitch , wavciength

Fig. 14.47 Susceptance versus pitch of wire grid in dielectric material of dielectric constant
of 1.15
(a) 0.01 X diameter wires
(b) 0 . 0 0 3 \ diameter wires
(c) 0.001 X diameter wires
0° incidence
180°
1180°
Radomes 513

0.1
02 0.3 0.1* 0.5 0.6 0.7

wire pitch,wavelengths

Fig. 14.48 Susceptance versus pitch of wire grid in dielectric material of dielectric constant
of 4
(a) 0.01 X diameter wires
(b) O.OO3X diameter wires
(c) O.OO1X diameter wires
0° incidence
180°
|| 80°
514 Radomes

(i)

c
o (ii)

So
20 30
frequency, GHz

(iii)

-1

•u

10 20 30 50

frequency, GHz
b
Fig. 14.49 Matching and loss of quartz fibre resin laminate (e = 3.15, tan 8 = 0.004, 1.7mm
thick) with wires
(a)
(i) 0° incidence dielectric only
(ii) compensated dielectric with wires
(iii) wires only, parallel to E vector
(b) angles of incidence
(i) 0°
(ii) 45° 1
(iii) 45° ||
Radomes 515
The A and B sandwich can, by choice of the core dielectric constant, give cover-
age at passbands which are not multiples of each other. The A sandwich can give
initial passbands in the ratio of more than two to one and the B less than two to
one. Fig. 14.51 shows several multi-band designs with skins of dielectric constant 4.
A core of e = 2 gives two passbands in the ratio 2.5/1, e = 3 gives 2.2/1 and
e = 6 gives 1.76/1. Increasing e to 10 provides three bands in the ratios of 2.5/1.6/1
(G, I and J bands.)

skin £ = 3.2
coil in core e = 1.1
skin e = 3.2

circumferential
wire matching in core

100

I 80
in
O

60

2f 3f
frequency
1 i

s0.8 •

0.6 -

2f 3f
frequency

Fig. 14.50 Loss and eHipticity versus 4:1 frequency band and over ± 30° antenna scan

14.7.3 Aberration and its reduction


The cause of aberration is asymmetry produced by the radome on the wavefront.
Rays passing asymmetrically through a radome, with different angles of incidence
and polarisation, will have different phase delays (insertion phase difference, IPD).
516 Radomes
A typical ogive-shaped radome with an antenna gimballed on axis is illustrated in
Fig. 14.52. With the antenna looking 10° off axis, say, through the single-layer
radome, the angles of incidence, polarisation, and the resultant insertion phase
difference (from Fig. 14.8), are as shown in Fig. 14.52. There is on average a greater
delay on one side of the aperture of approximately 10° phase. This asymmetry
results in a distortion of the phase front which would be expected to tilt the beam
towards the axis. A practical result for the aberration produced by this ogive
radome and a 17-wavelength diameter aperture antenna was 3.2 mr peaking near
10° scan angle, as shown in Fig. 14.52. The practical aberration curve is seldom
smooth, nor is it necessarily close to the theoretical calculation due to various fac-
tors, e.g. unquantifiable trapped and surface waves on the radome, and discontinu-
ity effects at its apex.

0-3 i
5S.2.2
0.2-

OG:OI:OJ
\ 1:10:25

0 0.1 0.2 0.3 0 0.1 0.2 0.3


skin thickness to free space wavelength d s /A

Fig. 14.51 Contours 100% transmission for lossless sandwiches, perpendicular polarisation,
O°, 50°, 70°, angles of incidence.
Angles of incidence: • # 0°; 50°; 70°
Dielectric constant: skin = 4; core = 2, 3, 6, and 10

Ideally, the phase delays for both parallel and perpendicular polarisations, irre-
spective of angle of incidence and wall thickness, should be identical. A study of
the phase-delay curves for dielectric panels, as exemplified in Figs. 14.8 and 14.17,
shows different and irregular phase delays, and hence a radome presenting a variety
of angles of incidences and/or different polarisations and/or operating over a band-
width will inevitably have aberrations.
Radomes 517

73°l 63°1

80°ll 60°il 62°n 60°ii 156° 131° 132° 131°

U3° 136° 131°

73°l 63°i U2° 132°

incident angles phase delays

with 17 X diam. antenna 10° scan in azimuth - E horizontal

- I* m radians
aberration
h3

10° azimuth scan,deg

Fig. 14.52 Aberration of single-layer ogive radome


Radius = 7 m, x = 2.5 m, / = 6.5 m, apex = 0.0
single layer e = 4, tan 6 = 0.015
thickness
= 0.28
wavelength
518 Radomes
The method of estimating the aberration caused by a radome has usually been
based on tracing rays from the antenna through the radome, including the effects of
multiple reflection, to determine the amplitude and phase in a plane external to the
radome. The tilt of the wavefront resulting from the summation of the rays enables
the aberration to be assessed. Improved methods have been developed,21'22 based
on calculating the near-field of the antenna and approximating the radome effects
by plane-wave transmission coefficients. In all cases, practical verification is usually
necessary to discover the actual aberration, and design modifications are often
required to obtain the best performance.
The shape of the radome clearly affects the aberration, and generally the more
pointed the radome, the larger is the aberration. Usually the shape is defined by
aerodynamic considerations, the material and constructional technique by structural
and environmental requirements, and the radome wall dimensions by the trans-
mission requirements. A study of the radome's transmission characteristics often
reveals that a design for optimum transmission does not necessarily produce mini-
mum aberration. A compromise design is often necessary: this involves varying the
thickness over the radome length.
An empirical guide to the interdependence of antenna size, radome shape, and
aberration slopes is given in Fig. 14.53, where the product of R, the sum of the
worst positive and negative aberration slopes, and antenna diameter d is plotted for
various fineness ratios. This data is based on solid half-wave radomes having dielec-
tric constants of about 4.
Aberration is one of the factors affecting the choice between single sheet and
sandwich designs. If, for instance, the single-sheet ogive radome of Fig. 14.52 is
replaced by an A sandwich design to save weight, the phase delays (from Fig.
14.11) would be as indicated in Fig. 14.54 for a 10° scan off axis. The phase delays
between each half of the antenna increase to 20° difference. The result is that this
sandwich radome, while saving weight, would be expected to give a higher aberra-
tion than the single sheet design. This is confirmed by the practical result shown in
Fig. 14.54.
If a very wide band is required, a single-layer design will be somewhat limited in
transmission performance, with large phase differences between parallel and perpen-
dicular polarisation. Under these conditions a sandwich design may be preferred, as
at least it can give better transmission properties.
The aberration characteristics may be modified by local changes to wall thick-
ness, such as added dielectric rings or patches23 which attempt to produce opposing
aberrations. These additions may be limited in effect, as they can mar the ampli-
tude distribution, thereby increasing sidelobe levels.
Metal elements can be used to minimise the phase difference between parallel
and perpendicular polarisations. Increasing the parallel phase delay to that of the
perpendicular can be achieved by adding extra skin thickness, but at the same time
inserting a grid of wires which, for perpendicular polarisation, provides an induct-
ance which compensates the capacitance of the added dielectric. An example is
shown in Fig. 14.55A. With increased skin thickness and wire compensation, the
Radomes 519
theoretical phase insertion delays for both polarisations can be made similar. A
practical result is shown in Fig. 14.55B.
Anisotropic dielectric materials24 can also be used to match the phase delays for

1-5 -i

1-0

AQ centre
wavelength

0-5

fineness ratio —

Fig. 14.53 Aberration slope and antenna size versus radome shape
L/D — fineness ratio
d/K = antenna aperture in wavelengths
R = sum of ± aberration slopes in degrees per degree

parallel and perpendicular polarisations. An example is shown in Fig. 14.56, where


the core has differing dielectric constants for the two polarisations. The resulting
reduction in aberration was similar to that obtained in Fig. 14.55.
520 Radomes

C
d

- 3
C
o

JO

15 10 10 15
azimuth scan,deg

Fig. 14.54 Aberration of an A sandwich ogive radome


a Ogive radome, A sandwich
Radius = 7 m , x = 2 . 5 m , y = 6 . 5 m , apex 0.0
thtckness
skins: e = 4; tan 5 O 0.015; = 0.03
wavelength
thit ness
core: e = 1.15;tan6 = 0.005; * = 0.3
wavelength
b phase delays through A sandwich antenna; 1 7 \ diameter at 10° azimuth scan
E horizontal
c Aberration
Radomes 521
14.7.4 Cross polarisation introduced by radomes, and its reduction
Cross polarisation introduced by the radome will degrade the performance of sys-
tems which require pure polarization, e.g. for reducing rain clutter, classifying
polarisation returns, frequency re-use satellites etc.

80 80 ~

V) 60
<u

O)

CD
a
a a.
JZ
Q.
20 20

40 60 80 40 60 80
angle of incidence, deg angle of incidence, deg

Fig. 14.55A Insertion phase difference of A sandwich with and without wires. Skins: e = 4,
tan 6 = 0.015. Core: e = 1.15, tan 6 = 0.005
a No wires
skins: 0.03X
core: 0.3\
b Wires: 0.008A. diameter
0.5\ spacing
outer skin: 0.03X
inner skin: 0.05\
core: 0.3\

14.7.4.1 Cross polarisation produced by a dielectric surface Consider a linearly


polarised wave E which is incident on a dielectric surface at angle /, with a polarisa-
tion angle 6 measured from the plane of incidence. (Fig. 14.57). The component of
parallel polarisation is EcosO, and of perpendicular is £sin0. On passage through
the radome the parallel polarised component will emerge with amplitude Tn(E cos6)
and phase <ph and the perpendicular polarised ray with amplitude Ti(Esin6) and
phase 02, where T and 0 are the amplitude and phase of the transmission coef-
ficients. These rays can be resolved into the main and cross-plane components.
522 Radomes

The sum of the main plane components is


Tm= sin20 02) 4-
and the sum of cross plane components is
Tc = Ey/Tjcos20 sin20 - cos20 sin20 cos0i - 2
jcos 0 sin

10 10 azimuth scar^deg

Fig. 14.55B Aberration of an A sandwich ogive radome wire compensated


Ogive radome
radius = 7 m.
apex = 0.0.
x = 2 . 5 m j = 6.5 m
dimensions as in Fig. 14.55A.
17\ diameter antenna
with wires
without wire compensation

so that the cross-polarisation ratio is


jTc(cross-plane)
Tm (main plane) cot20 + 2k 02) + k2 tan20
where k is the ratio of the amplitudes of the transmission coefficients for perpen-
dicular and parallel polarisation, TJT\\, and (0i — 0 2 ) is the phase difference for the
Radomes 523
100 H

90-

80-

70-

60-

r 50-

en

o
x:
QL

30-

20-

10-

-I 1 1 1 1 r- -i 1 1 r~
30 60 90

angle of incidence,deg
Fig. 14.56 Matched phase delays of A sandwich with anisotropic core.
perpendicular
parallel polarisation
thic ness
Skins: e = 4, tan 6 = 0.015, * = o.O3
wavelength

Core: eL = 1.3, tan 6 - 0.0005, t h i ° k n e S S . = 0.3


' wavelength
6,, = 1.5
524 Radomes

plane of
incidence

cross-polar plane

Fig. 14.57 Crosspolarisation through a dielectric


Radomes 525
particular angle of incidence, /. By adding all the main and cross-plane components
of all the rays from an antenna, the ratio of cross polarisation for the effective
aperture illumination can be determined, and the co-polar and cross-polar radiation
patterns can be calculated.25'26
For a circularly-polarised antenna, the incident components can be regarded as
E cos cot polarised parallel to the plane of incidence and E sin cot with perpendicular
polarisation. The transmitted components are then T$(E cos cot + <pt) and
Ti(E sin cot + <j)2). These can be resolved into left-hand and right-hand circularly-
polarised components with amplitudes A and B and associated phase angles at and
a2 such that
T\\(Eco$cot + 00 = A coscot + (*! + B coscot + a2
and
TL(E sin cot + <pt) = A sincot + ax — Bsincot + a2
Eliminating ax and ot2 from these equations gives the ratio of cross-polar to co-
polar signals as

A N 1 + k2 •
where k = TJT\\ as before.

14.7.4.2 Cross polarisation of a flat single layer and an A sandwich Fig. 14.58a
shows the cross polarisation for a flat single-layer radome of dielectric constant 4,
designed to be a half-wave thick at 0° incidence, for various angles of incidence and
polarisation angles, with linear polarisation incident. The cross polarisation
increases with angle of incidence and is greatest for a polarisation angle of 45°. Also
shown for comparison are the cross polarisation characteristics of an A sandwich
panel with skin and core dielectric constants of 4 and 1.2, and thicknesses of
0.08 cm and 0.9 cm, respectively, at I band. The A sandwich cross polarisation is a
few decibels greater than that of the solid half-wave radome. The behaviour of these
two panels with circular polarisation is shown in Fig. 14.58&. Also shown is the
improved performance obtained by modifying the thickness of the solid panel for
optimum performance at 60° incident angle.

14.7.4.3 Cross polarisation of curved radomes


(a) Spherical: The main and cross polarised diagrams of an antenna of 12 wave-
lengths aperture with a spherical single-layer radome are shown in Fig. 14.59, for
both linear and circular polarisations. The low levels obtained are due to the low
angles of incidence and the symmetry of the antenna within the radome. With an A
sandwich radome as in Section 14.7.4.2., the 45 and 50dB levels of cross polar-
isation of Fig. 14.59 would be increased to about 32 and 40 dB, respectively.
(b) Cylindrical: Fig. 14.60 shows the main and cross polarised patterns of an
antenna of 12 wavelengths aperture with a cylindrical half-wave solid radome, tilted
at various angles, for both linear and circular polarisation. With the A sandwich of
526 Radomes

CD
T3

10-
C
o

2
5 20

S
30-

30° 60° 90°


angle of incidence, deg

0° 30° 60° 90°


angle of incidence, deg

Fig. 14.58 Crosspolarisation of half-wave solid and A sandwich panels at angles of incidence
a Linear polarisation
A sandwich
A./2 solid at 0° incidence
e= 4
b A sandwich
(i) \ / 2 solid at 0° incidence
(ii) \ / 2 solid at 60° incidence
Radomes 527
Section 14.7.4.2., the crosspolarisation at 60° tilt increases to 20 dB for linear and
to 18 dB for circular polarisation.
(c) Conical: Main and cross polarised patterns for a conical half-wave radome with
the antenna on axis are shown in Fig. 14.61a, for linear and circular polarisations.
With the antenna off axis (Fig. 14.616), there is some lOdB increase in cross polar-
isation level. An A sandwich design as in Section 14.7.4.2 gives a general increase of
cross polarisation of 15 dB compared to the single layer half-wave design, as for
instance in the off axis antenna case illustrated in Fig. 14.61c.

50dB
v 3dB contours of
2dB main beam

Fig. 14.59 Crosspolarisation of spherical radome


a 12 in diameter dish of cosine distribution in solid \ / 2 radome of e = 4
b Linear polarisation
c Circular polarisation

(d) Ogive and paraboloidal: The diagrams are similar to the conical radome but tend
to give less cross polarisation, due to the averaging effect of different incident
angles.

14.7.4.4 Reduction of radome cross polarisation The degree of cross polarisa-


tion is dependent on the radome shape. Flat radomes can give large cross polarisa-
tion, particularly at the higher angles of incidence, with circular polarisation or if
the E vector is inclined to the plane of incidence for linear polarisation. Cylindrical
radomes also give large cross polarisation when the angles of incidence are high.
Sphere, cone, ogive and paraboloidal shaped radomes give little cross polarisation
with the antenna on axis but off-axis performance is generally worse than that of
flat radomes.
528 Radomes

The half-wave solid radome generally gives less cross polarisation than the A
sandwich, though off-frequency the solid can be worse than the A sandwich.
It is unfortunate that the radome cross polarisation effects are additive to the
'go' and 'return' path through the radome, thus the cross polarisation two-way is
6dB worse than one-way. For rain discrimination of 15 dB overall with circular
polarisation, the radome must introduce less than — 21 dB one-way with a perfect
antenna, and even less if the antenna is not perfect.

solid - radome e = 4

\ 12A diameter

azimuth ^ antenna with cosine


/ distribution

tilt in vertical plane

main beam

azimuth °° angle, deg5° azimuth °* angle, deg 5


*

Fig. 14.60 Crosspolarisation of cylindrical radome


a Solid \ / 2 radome, e = 4
b Linear polarisation
c Circular polarisation

Since in any practical measurement of ellipticity the radome and the antenna are
usually measured together, the actual performance of the radome can only be deter-
mined if the antenna alone is measured as well. In both cases the angle of the ellipse
as well as major to minor axis ratios must be measured so that the true contribution
from the radome alone can be established.

14.7.5 Reduction of radome reflections


Reflections from the radome surface back to the antenna can cause mismatch
effects, resulting in increased losses and ultimately in voltage breakdown.
Radomes 529

main beam
centre and outer circles -3 and -20dB levels

/ dB levels of
crosspolarisation

Fig. 14.61 Crosspolarisation values of a half-wave solid and an A sandwich conical radome
with antenna on and off axis, for circular and linear polarisation
a A solid conical radome with antenna on axis
b Half-wave solid radome with antenna off axis
c A sandwich conical radome with antenna off axis
(i) Linear polarisation, 0° scan
(ii) Circular polarisation, 0° scan
530 R adorn es

Reflections from the radome also degrade the antenna pattern and the tracking
accuracy. Typically, a reflected power level less than 1% or 2% is specified.
Amplitude and phase of reflections versus angle of incidence for both parallel
and perpendicular polarisation on various panels are shown in Figs. 14.62, 14.63,
and 14.64. All these designs have been chosen for minimum reflection near 70°
angle of incidence, as might be required for a high-speed nose radome.

22-

20- / -400

18- 1

;
16-
-300
V. 14-
-o

o
Z. 12- T
power r

-200

6-
-100
4-

2-

n
0 20 40 60 80
angle of incidence,deg

Fig. 14.62 Reflection from solid half-wave


e = 4, tan 6 = 0.015, 0.28X thick

The power levels reflected by the half-wave construction are less than 1% near
70° angles of incidence for both polarisations. Both increase at other angles, par-
ticularly for perpendicular polarisation. The A sandwich gives a minimum power
reflected near 70° angle of incidence for perpendicular polarisation, and near 63°
for parallel polarisation (Brewster angle). Its reflections increase considerably more
than those of the half-wave construction at other angles of incidence. The C
Radomes 531
sandwich reflection level, optimised near 70° angle of incidence, gives very low
levels of reflection over a wide range of angles of incidence, though for parallel
polarisation it increases at very high angles of incidence. The order of preference for
minimum reflection is generally the C sandwich, the half-wave solid, and lastly the
A sandwich, but if wide-band operation is required, the order could be altered.

400

300

-200

-100

20 L0 60 80
angle of incidence deg

Fig. 14.63 Reflection from A sandwich.


Skin: e = 4, tan 5 = 0.015, 0.03\ thick. Core e = 1.15, tan 6 = 0.005, 0.337X
thick.
Further improvement in reducing reflections from the C sandwich can be made
by choosing a higher dielectric constant for the central layer. Also the minima in
the reflection curves for perpendicular and parallel polarisations can be made to
occur at the same angle of incidence, by using an anisotropic core. Fig. 14.65 shows
an example of the performance with a higher dielectric constant centre layer of 5
(instead of 4), together with a core having a higher dielectric constant of 1.4
(instead of 1.35) for parallel polarisation.
532 Radomes

The reflections are less than 1% for both polarisations from 0° to 80° angle
of incidence.
If the dielectric constant is low, the reflections are smaller at low angles of inci-
dence, but can increase more rapidly at high angles of incidence for parallel polar-
isation, since the Brewster angle is lower. The choice of dielectric constant for mini-
mum reflection is thus dependent on the required range of angles of incidence.

20
1 - 400

16 11 1
- 300
I
i II

£12

1
power

1 - 200
1
8 1 ;

- j- ' --phase ^/
•^ > ^.
11 - 100

i
^ " \
/ y-
11 power
0
~ — • — .
/
T^— I

0 20 40 60 80
angles of incidence , deg

Fig. 14.64 Reflection from C sandwich


Skins: e = 4f tan 8 = 0.015, 0.036\ thick
Cores: e = 1.15, tan 6 = 0.005, 0.152\ thick
Centre: e = 4, tan 8 = 0.015, 0.072\ thick

Internal reflections in which rays reflected by the radome remain collimated will
produce subsidiary beams. This can be minimised by shaping the radome to avoid
linear surfaces (Fig. 14.66). Reflection from discontinuities can result in other
secondary radiation, as for instance at the apex of a nose radome.
Radomes 533
14,7,6 Effect of radome hardware (rain erosion cap, pitot systems and lightning
protection)
The foregoing discussion has been concerned with the dielectric transparency
effects on the antenna performance. Rain-erosion protective metallic caps, pitot-
static systems and lightning protection, which may be called for on airborne
radomes, will all contribute additional loss, reflection, aberration, pattern distor-
tion, side-lobe degradation and cross polarisation.

24-

20- 400

16
phase 300

V
12 \\
200

J-

-100

power

60 80
angles of incidence, deg

Fig. 14.65 Reflection of asymmetric C sandwich


Skins: e = 4, tan 6 = 0.015, 0.036\ thick
Cores: ei = 1.35, ejj = 1.4, tan 6 = 0.005, 0.15\ thick
Centre: e = 5, tan 6 = 0.015, 0.01X thick

(a) The rain erosion metallic cap A metal cap is usually placed on the nose of a
resin glass-fibre radome installed on a high-speed aircraft to protect the tip from
rain erosion. The effect of this aperture blockage can be represented by a new
source, similar in amplitude to the main antenna illumination, but in antiphase,
radiating with a pattern dependent on its dimensions. As the cap may be 3 cm in
534 Radomes

diameter its pattern at the longer wavelengths tends to be virtually omni-directional.


This subsidiary diagram interferes with the antenna's main beam and its sidelobes.
The effect is less significant when the ratio of the antenna diameter to that of the
cap is large. For example,27 when the diameter ratio is 10/1 the theoretical loss in
directivity is about 0.3 dB and the sidelobe degradation some 3 dB for a 24 dB side-
lobe. Though this serves as useful guide, the effect does not appear to be so damag-
ing in an actual radome situation, due to the discontinuity of the radome tip having
effectively caused a partial aperture blockage in any case. Additional aberration can
usually be detected, but in practice it does not appear to be comparable to that due
to the radome itself.

collimated and uncollimated


reflections

50
10 20 30 50
angle from rodome axis.deg

Fig. 14.66 Comparison of polar diagrams of antenna looking along axis of conical and ogive
radome
Radomes of half-wave solid dielectric e = 4, tan 6 = 0.015
£j_ to reflecting surfaces

(b) The pitot-static system When a pitot-static probe is mounted on a radome tip,
the rain erosion cap is not necessary, but since the mounting requires a larger di-
electric or metal support to be built into the radome nose, the aperture blockage is
typically at least 8 cm in diameter. Also due to the probe length, when the antenna
rays are not on the pitot axis, a shadow is produced which increases the blockage.
Radomes 535
The system requires three pipes, usually metallic, for pitot and static pressures and
probe heating, to run the length of the radome, and this causes further blockage.
Typical patterns for an antenna in an ogival radome are shown in Fig. 14.67. Loss
in directivity and increased sidelobes, particularly close to the main beam, are evi-
dent, together with a high reflection lobe from the metallic pipes. Aberration and
cross polarisation effects are greatly increased by this installation.

pitot static

antenna 15 X dia.

50

angle from radome axis.deg

Fig. 14.67 Effect of pitot installation on polar diagram of antenna looking along axis of
ogive radome (half-wave solid: e = 4, tan 5 = 0.015)
Ei to reflecting surfaces
ogive radome only
with pitot-static system

The effect on electrical performance may be so significant that the radome


designer will usually recommend a smaller-based shorter probe, or a pitot system
only, or else to move the system elsewhere. As the pitot-static system is usually
vital, not much compromise is possible, and perhaps the best that can be achieved is
a support designed with minimum blockage, with the pressure pipes made in a di-
electric material such as nylon. Electrical connections to the heater requires wires,
and although these can be of small gauge, they need to be protected against
536 Radomes
lightning strike damage, and in consequence are usually recommended to be con-
ained within a substantial copper pipe.
(c) Lightning strike protection Lightning strike protection may also be required to
prevent direct puncture of the radome wall, antenna damage etc. Conventionally,
this would lead to a series of conducting strips being placed on the outer surface of
the radome. The number and distribution of these strips is dependent upon the
degree of protection required and the electrical strength of the particular radome
constructions.28 The cross section of the conductor might range from a foil, which
could guaranteed to withstand one strike, to an area of the order of 10 mm2, which
will withstand multiple strikes.
Due to aerodynamic considerations, such strips would desirably be laid flat on
the radome surface and be about 10 mm wide. Their effect on microwave perform-
ance can therefore be considerable. An attempt has been made to reduce micro-
wave degradation by introducing an interrupted conductor, consisting of a line of
metal buttons each about 2.4 mm diameter, separated by a small gap. The onset of
the strike is said to lead progressively to the formation of an arc above the strip
surface. While offering an improvement in microwave performance, this type has
yet to be examined in detail to ensure its fitness for lightning protection.
Where it is considered that the internal metallic heater pipe of the pitot system
can also be used for lightning protection, the effect of collimated reflections can be
avoided by installing the pipe to follow an arc round the radome.

14.7.7 Prediction and reduction of radome hardware degradation


Prediction of microwave degradation caused by pitots and lightning protection
hardware is to date an inaccurate procedure. Ray tracing methods, in which rays
that intercept hardware are simply excluded, do not give a true representation.
Computational quality may be improved by exploitation of the Geometric theory
of diffraction (GTD), but further development is necessary. Work on a semi-
empirical approach,29 similar in concept to GTD, has been carried out as a first step
towards the prediction of conducting strip effects, in order to assess the effect of
strip dimensions, e.g. optimum aspect ratio for constant cross-section etc. Essen-
tially this work involved probing the field in the proximity of a strip in order to
assign an equivalent radiating line source which could subsequently be used in a pre-
diction programme.

14.8 Ground and shipborne radomes

A simple cover of dielectric material may be placed on an antenna to protect it


from the environment. For instance, small reflector-fed antennas may employ a
cover of dielectric material, either in sheet or sandwich form, with dimensions
chosen to give minimum transmission loss according to the data given in Section
14.3. The internal reflecting surface and the feed system is thereby protected from
the weather, but eradication of condensation may have to be considered. Simple
Radomes 537
solid covers may present such structural, weight and electrical problems, particu-
larly on larger rotable antennas, that they are not likely to be acceptable. A
stretched cover of fabric or plastic sheeting may reduce the weight problem, but
increases vulnerability to damage. In general, if large antenna systems are required
to have a covering, a separate radome system appears to be preferred.
Non-airborne radomes have been constructed in a variety of forms, such as air
supported, rigid foam, single skin, sandwich, and space frame. Various shapes and
sizes have been manufactured, dependent on the application. Fig. 14.68 shows an
example of a large space frame radome, whilst Fig. 14.3 illustrates some shipborne
radomes.

14.8.1 Environmental considerations for ground radomes


Ground and ship radomes have to withstand wind pressures, rain, snow, ice, light-
ning etc., and give satisfactory performance over possibly a — 50°C to 4- 80°C range
of temperature.

14.8.LI Wind pressure The pressure of air impinging at normal incidence on a


flat surface with velocity V is P- pV2/2, where p is the mass density of air. On
structures other than flat, the pressure distribution can be expressed asCp(0,0)P,
where 0 and 6 are angles from horizontal and vertical axes. CP will vary over the
surface of the radome, producing areas of positive and negative pressure (Fig. 14.41).
It can have a complex distribution dependent on shape, height above ground or
ship, and local objects and contours in the vicinity. Available wind pressure data is
usually given for a height of 10 m. At ground level, it may be 20% less and at 20 m
altitude some 10% more. On large projects, to obtain the most effective design,
wind-tunnel models tests are likely to be necessary.

14.8.1.2 Moisture and Rain effects on radomes The dielectric properties of


water with frequency are:—

Frequency Dielectric constant Loss tangent


1 Ghz 82 0.06
5 GHz 72 0.30
10 GHz 60 0.55
20 GHz 40 0.90

Moisture absorption by the radome will cause increased loss, since the high di-
electric constant of water will introduce high reflection losses, and the high loss
tangent will increase the aberration. It therefore is important that materials should
be impervious to moisture, and not deteriorate with time such that they become
absorbent.
Moisture accumulation and rain on the radome surfaces can produce similar
losses, which become increasingly serious at the shorter wavelengths, as the thick-
ness of water becomes more comparable to the wavelength. If moisture and rain
538 Radomes
form a film on the surface, Gibble30 has suggested the following formula to deter-
mine the thickness of the film: t = (3uRr/2w)1/3, where u and w are viscosity and
density of water. R is the radome radius, and r is the rainfall rate. Calculations of
the resultant attenuation at microwavelengths indicate somewhat higher values than

Fig. 14.68 Space-frame radome, Onsala, Sweden. Courtesy of Electronic Space Systems Cor-
poration

actually occur. A practical observation of a simulated 40 mm (1.57 in) per hour


rainfall rate showed that a water film thickness of 0.065 mm (0.0026 in) could be
obtained. The formation of such a thin film at D band, where X = 30 cm, will cause
little loss. At K band, where X = lcm, the electrical thickness of the water is sig-
nificant and some loss is expected. A film of water may not necessarily form, and
Radomes 539
much depends on the nature of radome surface, its shape, wind direction and force.
In practice, the Gibble formula does not agree with measured values for rigid metal
space-frame radomes, in particular those with special surface treatment of materials.
Under such conditions the rain does not form a uniform film, but water droplets
are formed, which run off as rivulets. They can be considered as covering a fraction
of the area of the radome surface and the extent of reflection and transmission loss
will depend on their effective area and depth. From data collected31'32 in practice
on various sizes of treated radomes, an effective thickness of water on the surface
of a rigid radome can be expressed by an empirical formula t = 0.48(QR)0'32, where
Q is the rain rate in inches per hour, R is the radius of the radome in feet, and t is
the thickness of water film in thousandths of an inch. This effective thickness is
shown in Fig. 14.69a for various rain rates and radome sizes. Smaller radomes have
a lower effective thickness and lower loss, since the collection area varies as the
diameter squared, but the circumference varies only as the diameter. A typical loss
versus frequency at 30mm/h rainfall rate on large space frame radomes is shown in
Fig. 14.696. In Fig. 14.69c, the loss is shown for various frequencies on large
radomes as a function of rainfall rate.
These results for treated surfaces indicate that at low frequencies, e.g. 1 GHz, the
loss tends to be insignificant, but becomes increasingly significant the higher the
frequency, particularly if the losses are doubled by two-way transmission and recep-
tion. The losses would be considerably greater if filming took place and the Gibble
formula applied. This emphasises the importance of surface treatment and material
selection to prevent filming.
On systems where noise temperature is important, the radome's contribution in
rain can be significant, but can again be minimised by choosing non-filming
surfaces. A 13m-diameter radome with non-filming surface in a 30mm/h rainfall
rate indicated an increase of only 5°K, whereas a filming surface would give a much
greater increase. On surfaces where water is absorbed, the increase could be an
order greater.
It is important to place these effects in perspective, and take into account the
additional attenuation of rain paths through the atmosphere (Fig. 14.70). At
1 GHz, the loss per kilometre is small, but at 5 GHz it begins to be significant. At
higher frequencies, the path loss tends to make the radome loss due to rain some-
what less significant. For instance, at 15 GHz, a path length of 4 km in 30mm/h
rain would give an attenuation of about 5 dB, which is greater than a treated
radome's loss of 1 dB (Fig. 14.69). However, it is obviously important to minimise
radome loss by choosing surface treatments which avoid filming and which dry off
rapidly.

14.8.1.3 Snow and ice and lightning protection of ground radomes The collec-
tion of snow depends on the shape of the radome. On spherical shapes, snow tends
to collect mainly on the top and lee side. It is often blown off, but, if not, it can be
removed by internal heating or pressure changes. Dry snow is not serious in contri-
buting to loss. For instance at 10 GHz, the dielectric constant is 1.25 and the loss
540 R adorn es

tangent about 0.004. This results in an attenuation of 0.25 dB, and about 5°K
increase in noise temperature, for a layer of 30 cm of dry snow. Wet snow can have
a considerable quantity of water present and tends to stick to most surfaces; and
consequently can result in significant attenuation, unless removed by heaters.

.c 80-
"E
E 60- 16m

2 20-

002 004
effective water thickness, mm

10 20 30
frequency, GHz

10 20 30 40
rain rate,, mm / h r

Fig. 14.69 Rain effects on ground radomes


a Rain rate versus water thickness and radome diameter
b Loss versus frequency w i t h 30 m m / h rain rate
c Loss versus rain rate and frequency

Ice films formed by freezing rain or moisture will give losses dependent on the
depth. Provided that no water is present, dry ice has a dielectric constant of 3.3 and
loss tangent of near 0.001, at I band. A build up of 1 cm depth, at 8 GHz, is likely
to give a loss of about 1.5 dB, and a noise temperature increase of 17°K. When ice
and wet snow removal is accomplished by internal heating, the surface should per-
mit it to slide off in the least time.
Radomes 541
Often the antenna is on an exposed site subject to lightning, and in this respect
the metallic space frame radome, properly earthed, behaves as a metallic protective
screen to the antenna. A purely dielectric radome can be punctured by lightning
unless some protection is incorporated.

2 5-

20-

1-5-

C
o
10-

0-5-

10 20
rain rate, mm/h

Fig. 14.70 Rain path attenuation versus rain rate and frequency

14.8.2 Air supported radomes


Some of the first ground-based radomes were air supported. A non-porous dielectric
fabric material was placed over the antenna system, sealed round the base, and then
inflated to clear the antenna. Rubberised fabric sheeting, as used in tents, was used
initially, as its loss tangent and thickness did not give significant transmission loss
at low frequencies. As frequencies of operation were increased, these losses became
more significant. Further, the material was prone to suffer from exposure, such that
moisture was absorbed, increasing losses, particularly at the shorter wavelengths
where water has a much higher loss tangent. Later, plastic materials such as Dacron
coated with Hypalon were found to give a better strength/weight ratio and better
weathering characteristics. Typical loss tangents in the microwave band for the
542 Radomes
rubberised fabric, as new, were 0.03 to 0.08, and for the later developed plastic
materials 0.005 to 0.02. Dielectric constants were approximately 3.5 for the
rubberised fabric and 3 for the plastic materials. The thickness depended on the size
of the radome, the strength of the material, and the environment it was likely to be
required to withstand. Typical thicknesses used were 0.5 to 1.5 mm.
The advantages of the air supported radome are:
(a) relative ease of erection and dismantling
(b) light weight and small packaging volume, easy for transport
(c) suitability for mobile stations
(d) relatively low cost
(e) good electromagnetic transparence and low reflection loss over wide frequency
bands
if) environmental protection not only for the antenna but for equipment and per-
sonnel
(g) more congenial for personnel and easier antenna system servicing.
Disadvantages are:
(a) requires continuous air pressure to maintain the support of the radome, and will
collapse with power failures
(b) can collapse due to fabric failure, due to abnormal wind or snow forces etc.
(c) if collapse occurs it can damage and even wreck the antenna system, particularly
if rotating
(d) mechanical and electrical properties tend to degrade with solar radiation,
moisture take up, contamination etc.
(e) requires considerable maintenance and inspection, with consequent out of
service time.
Inflatable radomes of up to 70 m diameter have been successfully used.33 Dual-
wall fabric inflated sandwich radomes have been used with air pressure between the
walls. These still suffer from the same disadvantages and usually have poorer elec-
trical characteristics than single skin radomes.

14.8.3 Rigid radomes


A rigid self-supporting type of radome construction is almost universally preferred,
so that the serious and often catastrophic collapse of the inflatable type is not
incurred. A further advantage of the fixed radome is that the antenna can be made
less robust, as it is never subject to the outside environmental forces, whereas the
inflatable radome could not permit this advantage to be taken. Further a high-speed
antenna requires less rotational power when inside a radome, as the internal air
tends to rotate as well, giving the advantage of less drag.

14.8.3.1 Rigid foam shell radomes Radomes, usually in the form of truncated
spheres, have been constructed of polyurethane foam. Foam panels can be jointed
with epoxy or polyurethane resin, such as to form a homogeneous low dielectric
Radomes 543
constant structure. The resulting radome has good transmission properties with very
wide bandwidths, particularly for the shorter wavelengths. The foam is required to
be moisture proof, and have a surface that permits rain to have low adhesion.
Examples of this type of radome are not common, and appear to be limited in
size.34

14.8.3.2 Rigid single skin radomes Panels, typically of resin glass-fibre laminate
construction with suitable flange attachments so that they can be jointed together,
have been used to construct radomes in numerous shapes, notably truncated
spheres. A limitation with this construction is that the larger the radome, the thicker
the skin has to be for structural reasons, and this limits the bandwidth and conse-
quently the operational versatility. For low reflection levels, the skin thickness
should preferably not exceed a twentieth of a wavelength in the dielectric material.
Reflections can be minimised by making the radome a half-wave thick or multiples
thereof, but only with bandwidth limitations.

14.8.3.3 Rigid sandwich radomes Small sandwich radomes have been made
without joints, but the larger types have usually employed shaped panels with
strengthened edge flanges to permit jointing. Large part-spherical radomes up to
45 m diameter have been constructed in this manner, with high transmission
efficiency, small boresight errors, and good performance up to 6 GHz.35 At higher
frequencies, the electrical performance falls off due to the joints in the panels and
to the increased thickness required for the skin for structural reasons. At the
shorter wavelengths, the quarter-wave thickness of the core of the A sandwich
becomes so thin, e.g. 3 mm at 25 GHz, that it presents structural problems which
generally preclude its usage.

14.8.3.4 Rigid space frame radomes The rigid space frame radome relies on the
frame for structural integrity. The windows are usually non-structural and of
relatively thin dielectric materials. In consequence, the electrical degradation of the
enclosed antenna is mainly determined by the aperture blocking of the frame and
the resulting induced currents, if metallic, and by the transmission and reflection
losses of the dielectric window materials.36
Such radomes are usually of cylindrical, spherical, or part spherical shape, and if
small, the frame structure may be sufficient if it consists of vertical dielectric or
metallic members, equally spaced round the circumference, and jointed together at
the top and base either at a point or to a ring. The effect on the electromagnetic
performance may be determined by model experiments, or by calculation of the
effect of individual frame members on the electromagnetic field and combining this
with the effect of the dielectric membrane windows. The aperture blocking of the
frame members will result in interference effects, introducing losses, sidelobe per-
turbations, and boresight errors, particularly if regularly spaced. With large diameter
radomes, the number of vertical members increases and for structural reasons hori-
zontal tie members have to be spaced round the radome to form a grid type space
544 Radomes

frame. Regular spacing of such horizontal and vertical elements tends to give rise to
a coherence of the scattered fields and maximises the deleterious effects on the
antenna pattern.
A considerable improvement in performance can be achieved by designing the
framework of random geometry such that reflections are incoherent ensuring much
smaller sidelobe perturbations, boresight errors, and an insensitivity to polarisation,
either linear or circular. Often the frame members are of metal rather than dielec-
tric, as they can usually be thinner, giving less aperture blockage. Typical radomes
with frameworks supporting dielectric windows are those of a quasi-random tri-
angulated construction, as shown in Fig. 14.71. These can be assembled on site to
the shape required, usually truncated spheres, taken apart for transport if required,
or lifted on or off the antenna as an entity as shown in Fig. 14.72. Large space
frame radomes have been constructed,15'37 mostly of a truncated spherical shape,
but occassionally toroidal, of up to 48 m diameter.38'39

Fig. 14.71 ESSCO radome undergoing installation


Courtesy of Electronic Space Systems Corporation

Structurally, the space frame consisting of a number of interconnecting beams


behaves as a shell, since the triangular sections are small compared to the whole
structure. For the stress analysis, the side conditions, particularly height and tem-
perature, must be known to evaluate the loads. The wind pressure is resisted mainly
by the membranes covering the frame, causing tensile stress resultants, which are
transferred to the frame. The beams of the frame can be subjected to radial and
Radomes 545
lateral buckling, i.e. inwards and sideways. Stresses due to various loads, such as
aerodynamic, gravity, ice, rain, earth movement etc., have to be considered in deter-
mining the structural design.40 Basically the important areas are the strength of the
panel membranes, their attachment to the beam framework, the stability of the
beams, their interconnections, and the structure as a whole.

Fig. 14.72 ESSCO 68 ft-diameter radome being lifted over communication antenna, Hawaii
Courtesy of Electronic Space Systems Corporation

Sangliolo41 emphasises the importance of the transfer of the loads through the
joint, where there is a potential of additional deflection above that which occurs
from the stretching of the skin and the frame, and has shown that increasing the
546 Radomes
joint stiffness and decreasing the membrane modulus will decrease the skin stress. A
nearly threefold reduction in the stress levels of the skin has been achieved with a
new membrane material, (Esscolam VI),15 which has a fourfold decrease in elastic
modulus. The material is non-absorbent and more flexible than conventional resin
glass-fibre laminate types of material and results in improved crack resistance. The
membrane's surface finish is of Tedlar, giving 90% rejection of solar radiation, thereby
helping to keep the enclosed antenna at an even temperature. The outer surface is
hydrophobic such that rain concentrates in beads and runs off quickly in rivulets.
The physical dimensions and dielectric properties of the struts and panels and
their distribution will have a direct effect on the electromagnetic performance of
the radome. A balance has to be struck between the size of the dielectric
membranes and the framework. The panel size should not be so large as to cause
the framework to produce a significant unbalance in the aperture blockage, for, as
the antenna scans this can produce excessive boresight shift, sidelobe deterioration,
and variable transmission coefficients. Neither should the panel size be so small as
to require such a large number of frames as to cause inordinate blockage. The
design of the frames, whatever their size and material, should consider their thick-
ness, shape and jointing so as to achieve the best electromagnetic performance.
A dielectric spaced frame has the advantage of having no fundamental frequency
cut off and gives the additional low-frequency coverage.
However, it has been found that reinforced plastic frames have more variable
characteristics and less suitable stress factors than metal frames and the latter are
invariably preferred for operation at frequencies above their electromagnetic cut
off. This situation could change with the possibility of more suitable dielectric
materials becoming available.
Over a wide range of frequencies, transmission loss will vary due to fundamen-
tally different mechanisms. As the windows of the metal spaced frame structure
become less than half-wave wide they will have a low frequency cut-off. Above this
low frequency cut-off, the transmission performance increases to a maximum, with
the metal framework tending to contribute a constant loss equivalent to that of the
physical blockage. At higher frequencies, the membrane losses become significant,
except at frequencies where the surface reflections cancel. Within the radome pass-
band the combination of the losses of the space frame and the windows are such as
to give, in typical designs, a maximum loss of ldB between 0.3 GHz and 30 GHz.
The loss is attributed to the physical blockage scattering some 5% of the energy,
and to the membrane giving 98% transmission at 0.3 GHz and 85% at 30 GHz. The
membrane loss is due to its thickness, typically 0.75 mm, its dielectric constant 2.8,
and its loss tangent 0.01, and the angle of incidence being near normal. At lower
frequencies, as the space frame approaches cut off (typically near 0.15 GHz) the
loss rises quickly. At higher frequencies, above 30 GHz, the loss fluctuates, as the
membrane becomes a half-wave or an odd quarter-wave thick, giving losses of 0.5
and 2 dB at the first pass and stop bands respectively with increased losses at even
higher frequencies. For minimum losses it is advantageous for the membrane mater-
ial to have a low dielectric constant and loss tangent, and be of minimum thickness.
Radomes 547
The space frame radome tends to give performance independent of polarisation,
due to the angle of incidence being near normal on the near-spherical surface, and
to the triangular frame members having a fairly even distribution of horizontal and
vertical blocking components.
Boresight errors will result from unsymmetrical disturbances of the phase front
of the antenna. They can be minimised by installing the antenna symmetrically
within the radome and for the radome to appear as symmetrical as possible to the
antenna over all scan angles. With the space frame radome, the large number of
metallic members ensures that the energy scattered from the frame structure,
typically 5%, is randomly polarised and distributed over wide angles. The boresight
shift errors are minimal, and with a 20m-diameter radome are found to be typically
less than 0.05 mr over the microwave pass band (3 to 30 GHz).
Sidelobe patterns of an antenna can be perturbed by the scattered energy from
the frame elements and windows of the radome. Minimisation of the perturbations
is achieved by the quasi-random structure and low reflections from the windows,
giving typically 5 to 10% non-coherent energy scattered over wide angles in both
planes. On a high-gain antenna, this scattered energy is likely to be at least at the
— 30dB level. A — 26 dB sidelobe of the antenna would be perturbed to a maxi-
mum of perhaps — 25 dB.
Noise temperature contributed by the radome is directly related to transmission
loss, which over most of the microwave band is not ohmic, but almost entirely
scattered energy. Most of the energy is scattered skywards, either directly or by the
antenna reflector, and only a small fraction reaches the 'hot' earth. Typically a 5°K
increase at 10 GHz is to be expected, decreasing at lower frequencies and increasing
at higher frequencies.
A summary of the main features of typical truncated-sphere metal space frame
radomes of diameter up to 50 m is:1S

Mechanical:
Antenna diam. Radome max. diam. Base diam. Height Weight
5m 7m 6m 6m 800 kg
10m 16m 14m 12m 5400 kg
20 m 33 m 30 m 12m 36000 kg
30 m 48 m 40 m 36 m 90000 kg
40 m 60 m 52 m 45 m 110000 kg
Frame elements: typical length 1.7 m of 6061-T62 aluminium
Membrane material: Tedlar bonded Esscolam VI, 0.95 mm typical thickness
Electrical: (with linear or circular polarisation)
Membrane material: dielectric constant 2.8, loss tangent 0.01
Transmission loss: 0.4 to 30 GHz, less than 1 dB
Sidelobe change: an increase of 1 dB on — 26 dB level
Noise temperature: an increase of 5°K at 10 GHz
548 Radomes
Bore sight shift: Radome diam. lGhz 3 GHz 10 GHz 30 GHz
(maximum) 5m 0.25 0.1 0.05 mr
10m 0.15 0.05 0.03 mr
20 m 0.17 0.05 0.03 0.02 mr
40 m 0.15 0.03 0.02 0.015 mr

-2 1

0.1 0.2 0.4 0.6 1. 4 6 10 20 40


frequency GHz

0.10 i r20 -?

0.2 0.4 0.6 1 4 6 10 20 40

frequency GHz
Fig. 14.73 Metal space frame performance
Radome 68 ft diameter; membrane 0.045 In; and space-frame geometry Essco
7000 series

The predicted performance of a 20 m-diameter radome, due to Essco,45 is shown


in Fig. 14.73, based on membrane loss, integrated frame calculations and model
measurements. Observed full-scale data on transmission losses, boresight shifts and
resultant noise temperature attributed to the radome confirm these calculations. On
small radomes, the boresight shift errors increase at low frequencies, since the
beamwidth is larger.
The replacement of earlier membrane materials, such as resin glass-fibre con-
structions, by improved dielectric materials, such as Esscolam VI,15 of lower
Radomes 549
dielectric constant, and better hydrophobic properties, together with developments
in space frame design have resulted in radomes capable of working in more severe
environmental conditions and at shorter wavelengths. Satisfactory performance has
been obtained with large reflectors operating as high as 100 GHz. This has only
been achieved with the reflector being enclosed in a temperature-controlled
environment and not subject to rain, snow, ice, solar heating etc. effects, thereby
permitting the reflector tolerances to be maintained.
Examples are the Brazilian 15 m telescope,42 which is enclosed in a 21m-
diameter metallic space frame radome designed for operating in the millimetre
band, and which gives a pointing accuracy of better than 20' of arc. The metallic
frame is designed to give less than 6% aperture blockage, and the membrane thick-
ness has been selected to give optimum transmission between 3 and 1 mm wave-
lengths, resulting in a total radome loss of less than 0.5 dB. Observations have been
successfully made at 85 GHz, and a similar-size telescope at Helsinki has confirmed
a similar performance at 80 GHz. The 20 m radio telescope at Onsala, Sweden43
(Fig. 14.71) is enclosed in a 30m-diameter, 24 m high radome and transmits
between 80 and 95% of the incident energy over the millimetre band, with observa-
tions taken as high as 110 GHz.
A 20m-diameter metallic space frame radome located at the Five College
Astronomy Observatory,44 University of Massachusetts, is designed for extended
use in the millimetre band. There is a problem at these wavelengths, in that there
are frequency bands where any water vapour in the air will result in absorption as
to preclude any observations. Further, at even shorter wavelengths, the absorption
tends to be near total. However, on cold days, when the water vapour is frozen, its
dielectric properties are such that absorption loss can be lower, and some observa-
tions may be made. Under these conditions, a large reflector antenna without
radome would not necessarily be able to make observations, since the thermal stres-
ses and the environment acting on an exposed antenna would cause significant
dimensional distortions, resulting in defocusing and gain losses.

14.9 Further radome developments

System performance requirements place increasing demands on radome design.


Specifications for ground radomes are to cover from very low frequencies to the
sub-millimetre band, and for airborne and missile radomes are to cover wider and
more bands simultaneously. There are requests for radomes for combined micro-
wave, laser and infra-red transmissions. At the same time the demand is for better
transmission and lower boresight shift, cross polarisation, sidelobe degradation and
reflection.
The designer's foremost demand to meet these requirements is for better
materials, which for space frame radomes will permit less blockage, better trans-
mission and lower noise temperatures, and for airborne radomes will permit higher-
temperature working, better rain erosion resistance, higher strength and be able to
provide a range of dielectric constants with low losses.
550 Radomes
The trend towards shorter operating wavelengths means that conventional
designs of half-wave thick solid, or sandwich constructions, become physically
thinner, usually introducing structural design problems and demands for new mater-
ials of greater strength. Wider and multi-band coverage requirements demand mater-
ials of low dielectric constant to give acceptable transmission properties.
While electrical designers are hampered by material limitations, the electromag-
netic understanding of radomes is far from complete. Particular problem areas are
aberration, trapped and surface waves, the effect of dielectric and metallic members
and hardware, and the reaction of radomes on antennas.
Computer-aided-design programs can speed up the structural and electromag-
netic analysis, but the assumption that all the input data is absolutely accurate is
often questionable. There is scope for the application of GTD in many areas which
may help solve special problems, particularly some of the discontinuity effects in
radomes. Near-field measurements could add to the knowledge of the behaviour of
rays received by antennas within radomes.
Radome production testing is often an expensive exercise, because of the large
number of electrical parameters, scan angles etc., that are required. Compact site
techniques currently being applied for testing antennas should be extended to their
associated radomes, to enable the radome contributions to pattern distortion to be
separated from those of the antenna.

14.10 Acknowledgements

The author wishes to acknowledge the kind assistance of those who gave permission
to publish photographs and data, and particularly to Dave Conti and Geoff Meades
of the Reinforced and Microwave Plastics Group, Stevenage, of British Aerospace
for their contributions to airborne radomes, and to Albert Cohen, President of
Essco, Concord, Mass., and his colleagues and users of their ground radomes, and to
A. M. Munro of RSRE, Ministry of Defence Procurement Executive, UK, for his
comments.

14.11 References

1 CADY, W., KARELITZ, M., and TURNER, L.: 'Radar scanners and radomes', MIT Radia-
tion Laboratory Series Vol. 26, McGraw-Hill, New York, 1948
2 HANSEN, R. C : 'Microwave scanning antennas', Academic Press, New York, 1964
3 WALTON, J. D.: 'Radome engineering handbook', Marcel Dekker, New York, 1976
4 SKOLNIK, M. L: 'Radar handbook', McGraw-Hill, New York, 1970
5 JASIK, H.: 'Antenna engineering handbook', McGraw-Hill, New York, 1961
6 'State of Radome Technology'. Proc. Symposia on Electromagnetic Windows, Georgia
Inst. of Technology, Atlanta, 1955-78
7 International Conferences on Electromagnetic Windows, Paris, 1967-75
8 MUNRO, A. M., VICKENDEN, B. V. A. et al\ 'Recent advances in radome design'. Con-
ference of Military Microwaves, London Oct. 1978
Radomes 551
9 BRYDON, G. M., and CARY, R. H.: 'Some radome materials with superior dielectric and
temperature characteristics'. First Int. Conf. of Electromagnetic Windows, Paris, 1967
10 EMERSON, X. and Cummings, X.: 'Data Sheet Eecospheres', Canton, Mass., USA
11 WALTON, J. D.: Ref. 3 Chap. 5 'Inorganic Radomes'
12 MOOREFIELD, S. A., and ROGERS, T. R.: 'The development of a lightweight advanced
aircraft radome'. Proc. Tenth Symp. on Electromagnetic Windows, Georgia Inst. of
Tech., 1970
13 CARY, R. H. J.: 'Results of dielectric and metallic loading of radomes'. Proc. Eleventh
Symp. on Electromagnetic Windows, Georgia Tech. 1971
14 WALTON, J. D.: Ref. 3, Chap. 8. FYALL, A. A.: 'Rain erosion - a special radome prob-
lem'
15 'Technical proposal for the Essco metal space frame radome'. Thiird Edition. Essco, Con-
cord, Mass., USA
16 Essco: Ref. 15 p. 27.
17 BERTRAM, H., BINI, P. and COOK, T.: 'Development of MRCA radome'. Third Int.
Conf. on Electromagnetic Windows, Paris, 1975
18 CARY, R. H.: 'Some novel techniques for avoiding obscurations', Radar-77. IEE Confer-
ence Pub.155
19 CARY,R. H.:Ref. 13
20 CARY, R. H.: 'Some design problems of an aircraft multi-band radome', Electromagnetic
Windo Symp. Georgia Tech., 1966
21 PARIS, D. T.: 'Computer aided radome analysis'. Proc. Tenth Symp. on Electromagnetic
Windows, Georgia Tech., 1970
22 DELFOUR, A., DUPUTZ, J. F., and SARREMEJEAN, A.: 'New method of calculation of
aberration'. Third Int. Conf. Electromagnetic Windows, Paris, 1975
23 SELLER, D. C : 'Aberration of dielectric patches and rings'. First Int. COnf. on Electro-
magnetic Windows, 1967
24 MUNRO, A.M., WICKENDEN, B. V. A., and WARD, F. S.: 'Artificial and anisotropic
materials'. Third Int. Conf. on Electromagnetic Windows, Paris, 1975
25 DOWSETT, P.: 'Cross polarisation in radomes: A program for its computation', IEEE
Trans. AES9, p. 42
26 CARY, R. H.: 'Cross polarisation produced by radomes', Aerospace Antennas, IEE Conf.
Pub. 77,1971
27 'Engineers handbook', Micro wave J. Feb. 1968, p. 133
28 CARY, R. H., and CONTI, D. A.: 'The protection of aircraft radomes against lightning
strike'. Proc. 13th Symp. Electromagnetic Windows, Georgia Tech., June 1976
29 CARY, R. H., and CONTI, D. A.: 'Radome obscurations and their equivalent sources',
Proc. 12th Symp. Electromagnetic Windows, Georgia Tech., June 1974
30 GIBBLE, D.: 'Effect of rain on transmission performance of a satellite communication
system'. IEEE Int. Convention Record, Part VI, March 1964
31 HENSEL, S. L., and SMOLSKI, A. P.: 'Performance of Essco metal space frame radomes
during precipitation'. Essco, Concord, Mass., Report MTP-77-0011
32 COHEN, A., and SMOLSKI, A. P.: 'Effect of rain on satellite communication earth
terminals and rigid radomes', Microwave J. Sept. 1966
33 BIRD, W. W.: 'Large air supported radomes for satellite communication ground stations'.
Proc. OSU-RTD Symp. Electromagnetic Windows, June 1964
34 LAVRENCH, W.: 'Polyurethane foam radome tests at C Band'. Nat. Res. Council Canada,
ERB-5 33 July 1960
35 CURTIS, R. B.: 'Survey of ground radomes'. Rome Air Dev. Center RADC-TDR-64-127
36 KAY, A. L.: 'Electrical design of metal space frame radomes', IEEE Trans., AP13, March
1965
37 'Technical data on five types of ground radomes made by Goodyear Aircraft Coy'. GER
10064 Nov. 1960
552 Radomes
38 COHEN, A., DAVIS, P., NILO, S. C, and ORABONA, J. F;: 4A 150 ft metal space
radome', Proc. OSU-WALDC Radome Symp. WADC Tech. Rpt. 57-314
39 WEISS, H. G.: 'The Haystack experimental facility', MIT Lincoln Lab. Tech.Rpt. 365,
Sept. 1965
40 D'AMATO, R.: 'Metal space frame radome design', Int. Symp. Structures Technology for
Large Radio and Radar Telescope Systems, MIT, Oct. 1967.
41 SANGIOLO, J. B., and ROHWER, A. B.: 'Structural design improvements of Essco
radomes and antennas'. IEEE MERS-77, Washington, DC, Nov. 1977
42 KAUFMANN, P., and D'AMATO, R.: 'A Brazilian radio telescope for millimetre wave-
lengths', Sky and Telescope, 45, March 1973
43 MENZEL, D. H;: 'A new radio telescope for Sweden', Sky and Telescope, 52, Oct. 1976
44 ARNY, T., and VALERIANI, G.: 'The Five College radio astronomy observatory', Sky and
Telescope, 53, June 1977
45 ESSCO: Ref. 15, p. 7
Chapter 15

VLF, LF and MF antennas


John S. Bel rose

15.1 Introduction

The frequency band of interest in this Chapter, 10 kHz to 3 MHz, extends over a
very large range of wavelengths, from 30 km to 100 m. This means that realisable
structures are small with respect to the radio wavelength at VLF frequencies,
whereas at MF the electrical height of antennas employed can be one-half wave-
length or more. Typical tower heights employed at VLF are 300 m, which at
15 kHz are only 5.4 electrical degrees in height. Equivalent tower heights at 1500
kHz would be only 3 m high. Towers of heights up to 300 m are sometimes
employed at MF; however, since such a tower would be 360° high at 1000 kHz, it
would be sectionalised with each half being fed in phase so that its directivity
would be towards the horizon. Propagation in this frequency range is dominated by
vertical polarisation (ground and skywaves), cf. Belrose11, and so the object in
medium- and long-wave transmission is to radiate vertical polarisation, and to
provide as great a field strength as possible in a horizontal plane.
Since antenna dimensions and frequencies scale in inverse proportion, the
antennas discussed are not unique to the lower frequency bands. In fact some
of the results presented in this Chapter were measured at scale model frequencies
on an antenna pattern range. The same antenna fundamentals apply at any fre-
quency, but with a few important differences. At the higher frequencies, a quarter-
wavelength or a half-wavelength becomes a practical dimension, and at HF, VHF
and UHF monopole antennas can be elevated, remote from the ground beneath.
The radiation pattern for vertical elevation angles below the horizon must therefore
be considered, and the 3 dB gain attributed to ground-based monopole antennas,
because they radiate over a hemisphere rather than a sphere does not apply. Also,
ground loss resistance does not scale directly with frequency. The emphasis in this
Chapter is on vertical antennas, or antennas that radiate dominantly vertical polar-
ization, for use at the lower frequencies, and therefore discussion about ground
losses, radiation efficiency, etc. refer to this application.
554 VLF, LF and MF antennas

15.2 Analysis

15,2.1 Characteristics of vertical antennas


If the vertical dipole is some distance above the earth, its impedance and radiation
characteristics are identical with a free-space dipole. However, vertical antennas at
VLF, LF and MF are usually not far removed from the ground, and generally the
antenna is fed against the ground plane as a monopole radiator. Such an antenna
is said to be base-fed or series-fed, or base-loaded in the case of electrically short
antennas. For the purposes of analysis, neglecting losses in the ground, the mono-
pole can be considered as a dipole whose lower half has been cut off and replaced
by its image in the ground. For some purposes, the concept of the monopole
radiator and its image is useful. The currents on the vertical parts of the radiator
and on its image flow in the same direction, see Fig. 15.1; i.e. the currents are in
phase. The currents on the horizontal parts of the radiator and its image are out
of phase. If the height of the antenna is small compared with the wavelength,
radiation due to current flowing on the horizontal part of the antenna is almost
exactly cancelled by that due to the oppositely phased current on its image, and
the antenna is described as having a 'non-radiating' top load. The equivalent length
of the monopole, and its input impedance, is one-half that of the corresponding
dipole antenna. The power gain of the 'grounded' monopole antenna is, however,
3 dB greater that a dipole antenna in free space, see Fig. 15.2. This is because, for
a grounded antenna, all the power is radiated into a hemisphere, rather than a
sphere.

. — e)

Fig. 15.1 Current distribution on typical monopole antennas and their images

The sketches shown in Fig. 15.1 (except d) illustrate the antenna as though it
were grounded, and such a radiator is usually described as being a 'grounded-
antenna' to differentiate it from the ungrounded antenna shown in Fig. 15.Id.
VLF, LF and MF antennas 555

Actually, the antenna is not directly grounded, but it is resonated and matched,
and fed against ground through the output impedance of the transmitter. In some
circumstances, the base of the antenna is indeed grounded and the antenna is not
base-fed, but it is fed at some distance above the ground plane. This arrangement
not only eliminates the requirement for a base insulator, but it can alter the current
distribution on the radiator, and hence radiation from it.

mV/m
TYPE OF CURRENT PATTERN AT I km dBd
ANTENNA DISTRIBUTION SHAPE Eo GAIN

ELECTRICALLY
SHORT ANTENNA

"ftySf /I 1 - ^ MO 4 . . .

39
ANTENNA 1 I — ^ « 50'

ATOM*

M A I C UU/VV/CT
ANTENNA CENTRE V \/J 456 6.27
X/Z ABOVE GROUNO

TWO END ON
HALF WAVE 1 ««7
53?
? p.a
7 68
IN PHASE ANTENNAS » ' '
{FRANKLIN ANTENNA)

Fig. 15.2 Current distribution, vertical pattern shape, field strength and gain, on the assump-
tion of sinusoidal distribution of currents and a perfectly conducting ground, for
monopole antennas of various heights. The gain is referred to a dipole antenna
in free space

At VLF and LF, practical antennas are electrically short, and, as we shall discuss
below, to achieve reasonable radiation efficiencies and input impedances, they are
usually top-loaded to produce as nearly as possible uniform current on the radiator.
The radiation patterns for such antennas closely approximate that for a hypotheti-
cal vertical current element; for a lossless electrically short antenna, the field
strength at 1 km for 1 kW of transmitter power is 300 mV/m.
As the frequency is increased, for a vertical antenna having a particular physical
length, the radiation resistance referred to the base of the antenna increases, the
antenna reactance decreases and the radiation efficiency increases. At the lowest
frequency for which the input reactance is zero (this is the fundamental frequency
of the antenna system) the antenna is one-quarter wavelength long or 90° in electri-
cal height. The field strength at the horizon, see Fig. 15.2, is slightly greater over
that for the vertical current element.
556 VLF, LFand MF antennas
These facts are well known, since the publication of two historic papers by
Ballantine6'7 in 1924 led to the present development of vertical antennas for MF
broadcasting. In these papers, it was shown that for vertical antennas higher than
one-quarter wavelength, the radiation resistance referred to the base continued to
rise and went through a very high value when the height of the radiator approxi-
mated to one-half a wavelength. This then pointed to a method of increasing the
radiation efficiency by using antennas having a radiation resistance very large with
respect to the ground loss resistance. Since the antenna pattern is more squashed
towards the horizon, its gain is increased (see Fig. 15.2). The second paper dis-
closed that there was an optimum height for a grounded vertical antenna, a height
of 0.64 wavelength or 231°. The gain of a 0.64A radiator is 6.03 dBd (dB gain
relative to a dipole antenna in free space). This is the maximum gain for a linear
grounded antenna. Increasing the height beyond this optimum value results in a
decrease in the field strength measured at the horizon in front of the antenna, and
an increase in the high-angle lobe. When the antenna is X in height, the low-angle
lobe has decreased to zero, and there is only the high-angle lobe.
The field strengths and gain figures quoted apply to an idealised infinitely thin
wire antenna, along which current propagates with the velocity of light. For practi-
cal antennas, the velocity of propagation is reduced, as is the maximum gain (cf.
Surutka and Popovic90).
Diligent research and experiments have been conducted for other possible cur-
rent distributions (cf. Brown20) that might equal or surpass the optimum design
discussed by Ballantine, but only two types, the half-wave radiator elevated A/4
above the ground plane and the Uniphase antenna, surpass the straight vertical
antenna whose height is about 5X/8. For the elevated half-wave radiator, both the
mutual impedance betweeen the radiator and its image, and the phases between
the direct and ground reflected waves are optimum, so that they combine and
provide a maximum gain (6.27 dBd). The gain of two end-on half-wave in-phase
antennas (a simple Franklin type antenna) is 7.68 dBd, and this radiator has only
a small high-angle lobe. Antennas having greater vertical extent become imprac-
tical for the frequency range of interest. At MF, multi-feed is sometimes employed,
to control more accurately the current distribution on the radiator and hence its
vertical pattern.

15.2.2 Parameters needed to specify performance


Antenna design depends on various considerations:
(a) power radiated P r , by the antenna, which is usually expressed in terms of
field strength at ground level in front of the antenna in the direction of interest, or
in terms of the radiation efficiency of the antenna. It is usual to specify the un-
attenuated field strength at 1 km for 1 kW of input power. That is, the effect due to
the finite conductivity of the ground on the propagation of the ground wave has
been eliminated, so that the radiation efficiency of the antenna can be compared
with other antennas or antenna types installed over ground of differing conductivity;
(b) the input impedance Za, which is the input impedance at the driving
VL F, LF and MF an tennas 557
point. If the antenna is base-fed, we refer to its input impedance as being Zb =
Rb ±jXb ohms. This is the impedance which must be tuned and matched to the
transmission line feeding the antenna and to the transmitter output impedance.
The antenna resistance Rb is the sum of the radiation resistance Rri the effective
ground loss resistance Rg referenced to the feed point, and conductor and effective
insulator loss resistance if these are not negligible. The resistance Rc of the tuning
coil which is needed to resonate electrically short antennas must also be included in
the calculation of radiation efficiencies since this can be a dominant loss resistance;
(c) the antenna bandwidth;
(d) the power limitations for the antenna; and
(e) the antenna radiation patterns.
For VLF, the azimuthal pattern of the transmitting antenna is generally omnidirec-
tional, but at MF both the azimuthal and vertical radiation patterns are important
and can be optimised for the transmission requirements. Directional antennas are
employed for MF broadcasting, to protect a listener receiving the same frequency
in an adjacent broadcast area, since frequency sharing is a broadcast requirement,
especially in North America. The vertical pattern gives a measure of the relative field
strengths between ground and sky waves, which are important for estimating the fade-
free radius of a broadcast station, particularly at night, when sky waves are strong.
None of the above characteristics can be determined without a knowledge of
the current distribution on the radiator, which is usually assumed to be sinusoidal
for the usual methods of analysis. It can, however, be calculated, along with other
antenna design parameters, if a more complicated numerical moment method is
employed for analysis.
The following Sections consider the methods of analysis, and the application of
these methods for particular basic antenna types. In practice, differences between
theory and/or prediction by model measurements may arise in full-scale appli-
cation, because of the effect of supporting guys, base capacitance to ground, earth
losses, re-radiation from adjacent structures (e.g. other antenna towers, power
lines, highrise buildings), terrain effects etc. Each of these are considered briefly in the
following Sections, first in overview, and afterwards in more detail for specific antenna
designs. Naturally full detail cannot be given; this must be obtained from the material
referenced if the antenna has been studied in sufficient detail, although there remains
much still to be done and so the final explanation cannot always be given.

15.2.3 Methods for calculating antenna parameters


There is no great difficulty, at least in principle, in determining the radiation field
produced by an arbitrary current distribution. It is when we are given a metallic
antenna and an impressed EMF, either localised or distributed, that the difficulty
arises. In this case, the current distribution is yet to be determined, and that is not
easy, before the antenna can be analysed by one or more of the standard methods.
Fortunately, for the simpler antennas, a sinusoidal current distribution is a fair
approximation, and the results obtained assuming such a current distribution give
satisfactory agreement with practical measurements. A major problem is that of
558 VLF, LF and MF antennas

determining the effect of the finitely conducting ground on the terminal impedance
of the antenna, and this is especially difficult from a theoretical point of view when
the antenna is fed against a ground screen, the 'connection' to the ground being
through this radial-wire ground system.
The finitely conducting ground also influences the vertical polar diagram, but
this effect is somewhat easier to deduce theoretically. It is the ground many wave-
lengths in front of the antenna, where only the radiation fields are significant (the
electrostatic and induction fields are negligible), that affect the vertical pattern.
Various approaches to the antenna problem are: (a) the Poynting vector method;
(Z?) the induced EMF method; (c) the equivalent transmission line method; and
(d) the numerical moment method of analysis.
In the induced-EMF method, the electric field produced parallel to the antenna
by a known or an assumed current distribution is calculated and used to compute
the distributed EMF along the length of the antenna. By taking the product of this
EMF and the current at every point, and integrating over the length of the antenna,
the complex flow of power out of the antenna is obtained. The complex impedance
at every point along the antenna is also obtained directly. Since the computations
deal with the near fields of the antenna, the method can also be used to calculate
the mutual impedances between antennas (Jordan and Balmain46), or between
re-radiating structures that affect the pattern of the antenna. The method is very
suitable for computer analysis. It is, however, only useful for 'thin' structures.

rs
JL

(a) (b) (c)

Fig. 15.3 An antenna as an opened-out transmission line

A simpler, yet practical approach which gives answers with sufficient accuracy
for engineering design, is based on the analogy of an open-circuited transmission
line, see Fig. 15.3. On the assumption of a sinusoidal current distribution, we can
calculate the antenna reactance, and the radiation resistance. If greater accuracy
is required, a closer approximation to the current distribution can be obtained
by assuming a dissipative transmission line, whose losses just account for the
radiation losses. A further refinement is to include some estimated value of dead-
loss resistance, i.e. earth losses etc., in the evaluation of the equivalent dissipative
transmission line. This method is used extensively in this Chapter.
VLF, LF and MF antennas 559
The most rigorous method is based on a numerical moment method of analysis.
The basic approach to the solution of the antenna integral equation it to rewrite
it as a set of linear algebraic equations, in which the unknown quantities are a
suitable expansion of the current distribution. After dividing the conducting
antenna wire or tower into subsections, the unknown current on each is related to
the incident field by an impedance matrix. Inversion of this matrix gives an admit-
tance matrix, that allows the current distribution to be found for any source field.
Although not the first to use matrix solutions, there is little doubt that Harrington's38
unified treatment of matrix methods for field problems marked a turning point in
the direction of antenna analysis (Ramsdale71). Since then, numerical techniques
have been used for a wide variety of antenna problems. Numerical methods used to
calculate some of the results given in this chapter employed NEC (numerical
electromagnetic code), developed by Burke.24'25 The calculations were, in general,
carried out for dipole antennas and re-radiating structures twice their actual size.
Employing the methods of images, the results obtained therefore correspond to
monopole antennas over a perfectly conducting ground plane. The inclusion of a
finitely conducting ground beneath the antenna and its radial ground system has
also been numerically modelled by Burke et al, 26 but their solution is only rigor-
ously correct if the antenna system does not make physical connection with the
ground plane. This will be considered later.

15.2.4 Radiation resistance


When the radiated power is known, it can be expressed in terms of a radiation
resistance and a current at some point on the antenna, usually where the current
is a maximum. For antennas shorter than A/4, the maximum current is at the base
of the antenna and the radiation resistance is then referenced to that point. When
the antenna height h is greater than A/4, the maximum current occurs at some
point above the base of the antenna. In this circumstance, the radiation resistance
is usually referenced to this point, which is known as the current loop. The reason
for this is obvious. Since by definition

Pr=Ia2Rr watts (15.1)


Pr = power radiated
Rr = radiation resistance

there would be great uncertainty in calculating the radiated power using this
expression at a point on the antenna where the current is small or near zero.
In the case of electrically short antennas, the radiation resistance is

Z? r = 10G 2 ohms (15.2)


where G is the electrical height of the antenna in radians. If the base current Ib is
taken as 1.0 A, for comparative purposes (following Laport56), this relation can be
transformed into a very useful form that shows that the radiation resistance is a
560 VLF, LF and MF antennas

function of the degree-ampere area A of the plot of current distribution on the


radiator. This is
Rr = 0m2SA2 ohms (15.3)
The utility of this relation is that it holds for any shape of in-phase current on the
vertical parts of the antenna, and it can be used to calculate the radiation resistance
with reasonable accuracy for straight vertical radiators for heights up to about 0.2X,
and as well for top-loaded and inductively-loaded antennas.
For a straight vertical antenna with no top-loading, having an assumed sinusoidal
current distribution on it
180 l - c o s G Q ^
A = degree-amperes (15.4)
77 sin Ga
where Ga = effective electrical height of the antenna.
Electrically short antennas (Ga < 30°) have a small radiation resistance and a
relatively high capacitive reactance. Since, for practical antennas, the ground loss
resistance is not negligible with respect to the radiation resistance, some form of
top-loading must be employed to increase the current area on the vertical part of
the antenna, and hence to increase the radiation resistance and reduce the antenna
reactance. This increases the bandwidth of the antenna and the radiation efficiency.
For a top-loaded antenna the current area on the vertical part of the antenna
can be calculated from the expression
180 cos 1(G2a-Gv)-cos Ga j
A = - degree-amperes (15.5)
7T sin Ga
where Ga = electrical height of the entire antenna
Gv = equivalent electrical height of the vertical portion of the radiator

The electrical height of the entire antenna at its operating frequency/is

Ga = y - (90°) degrees

where f0 is the resonant frequency of the entire antenna, either calculated or


measured.
In the case of an antenna with top-loading, where the top-loading is adjusted so
the antenna is resonant (Ga = 90°) at the operating frequency (f = f0), the current
area on the antenna is for practical purposes a maximum, and is equal to
180
Ao = sin Gv degree-amperes (15.6)
n
The maximum current area Amax occurs for G a > 9 0 ° , in fact when Ga =
90 + Gv/2 degrees, and

Amax ~ (360/7T) sin Gv/2 (2 — cos Gv/2) degree-amperes (15.7)


VLF, LF and MF an tennas 561

However, for Gv < 20°, Rr (Ga = 90°) is within 3% of Rr(max) and at LF and
below, when Ga is slightly greater than 90°, the input impedance is low and
inductive; hence difficult to tune.

11 o—
^ ^
/
§90- —
\
/ \
70-
/
/
50-
/

30-
/

60 80 I00 I20 I40 I60 I8O 200 220


ELECTRICAL HEIGHT (G) DEGREES

Fig. 15.4 Radiation resistance of a monopole antenna referred to the point on the antenna
where the current is a maximum, as a function of the equivalent electrical height
of the antenna

When the height of a straight vertical radiator is greater than about 0.2X (72°),
a more exact expression, based on the induced EMF method, should be used to
calculate the radiation resistance. In this case Rr (loop), the radiation resistance
referred to the loop current, may be obtained with good accuracy from the approxi-
mation (Jordan and Balmain46)

i? r (loop) = 15 | - ~ sin 2Ga + ( I n - ~ - + 1.7221cos 2Gn + 4.83


2 \ 180
+ 2 In ohms (15.8)
180
where, as before, Ga is the electrical height of the antenna. The relation between
i? r (loop) and i? r (base) and between the equivalent electrical height Ga (degrees)
of the vertical antenna and its physical height h in metres or H in degrees will be
considered in a following Section. For the moment, it should be noted that the
usual relation Rr (base) = Rr (loop)/sin2 Ga should not be used for tower antennas,
since the actual base resistance can differ widely from this expression, which is only
accurate for very thin antennas. Values of Z?r(loop) versus antenna height Ga
(in degrees) calculated from eqn. 15.8 for thin antennas are graphed in Fig.
15.4.
562 VLF, LF and MF antennas

15.2.5 Antenna impedance


To a great extent, a vertical antenna behaves like a transmission line, which is open-
circuited in the case of a straight vertical antenna, or terminated in the reactance of
the top-loading for electrically short capacitively-loaded antennas. The analogy of a
vertical antenna as an opened-out transmission line was briefly discussed above,
and illustrated in Fig. 15.3.
Although the average characteristic impedance of the antenna cannot be defined
exactly, it appears to be sufficiently well established for practical purposes. Of the
several formulae put forward, cf. Howe,42 Labus53 and Sigel and Labus,82'83
Schelkunoff81 and Laport,56 the only one that is consistent with measurement is
that of Howe, which is

Zo = 6oAn£-ij (15.9)
where h = height of radiator
a = radius of radiator
In the derivation of this formula, it was assumed that the antenna had a cylindri-
cal cross-section. However, it can be used for structures having other cross-sections
by application of an equivalent radius. Since the replacement of a complex tower
shape by a cylindrical antenna conductor is not an exact one, it is expected that no
precise relation exists between the conductor radius a and a tower of artibrary
cross-section. Hallen,36 Lo60 and others have calculated the equivalent radius of a
polygon by the in variance of the capacitance per unit length, and, for a triangular
and a square tower of side width b, they calculate that a is equal to 0.42b and
0.59&, respectively.
Jaggard43 describes the application of isoperimetric inequalities to calculate
equivalent radii, a method which can be employed for structures of any shape.
Here we bound and approximate the equivalent radius aeq by the following
inequality:
i

< Per/2 n (15.10)

where Area is the area of the tower, and Per is the perimeter. Taking the arithmetic
mean of the above bounds yields an approximation to the equivalent radius, which
for the simple polygonal structures that have been analysed is as accurate as the
more exact solution.
Applying this transmission-line analogy, the reactance of an open-circuited
transmission line of characteristic impedance Zo and electrical height Ga is
Xh = -jZS cot Ga ohms (15.11)
With capacitive top-loading of a vertical radiator of electrical height Gv and charac-
teristic impedance Z%9 with a flat-top reactance X at a frequency/, the reactance at
the feedpoint would be (see Laport56)
VLF, LF and MF antennas 563

_ XcotGv+iZSmiGv
Xb Zo (15J2)
- ZZcosGv+jXsinGv
Both X and Gv are functions of frequency, so that computations can be made for
all values of / . At the fundamental frequency f0 of the antenna, Xb = 0, under
which condition the numerator of this equation must be zero. The terminating
reactance for resonance X = Xo is:
Xo (at/o) = -j'ZS tan Gv (15.13)
An alternative and perhaps simpler method of calculating the base reactance is
to equate the top-load reactance X to an equivalent lengthening of the antenna
(0) of characteristic impedance ZJ, given by
X = ZS cot 6 (15.14)
Since X has been calculated, we can solve eqn. 15.14 for 0, and the base reactance
is then
Xb = -j2%cot(Gv + 0) (15.15)
These equations strictly apply for electrically short antennas (h < 0.2X).
For straight vertical antennas more than 0.2X high, the more exact expressions for
low-loss but not loss-less transmission lines must be employed. In the transmission-
line representation, the radiated power is replaced by an equal amount of power
dissipated as ohmic loss along the line. This power loss is assumed to be due to a
series resistance r ohms per unit length, and a shunt conductance g mhos per unit
length. Furthermore, it is assumed that the I2r loss per unit length at a current
maximum is equal to the V^g loss at a voltage maximum. This approach to the
antenna problem was first given by Siegel and Lab us,83 and with slight modification,
the essentials of the method are detailed in Jordan and Balmain.46 Their application
of the method, however, gives incorrect results due in part to the fact that they
employ a formulation for the characteristic impedance of the antenna which is not in
accord with experiment, but in the main to the fact that they incorrectly estimate
the electrical height G of the monopole from the physical dimensions of the tower
height H and effective tower diameter D expressed in electrical degrees. Modified
expressions for the impedance of antennas calculated on the transmission line
method, for h > 0.2A follow:

2 <1516>
coshrT^l-cos Ga
Zn

- fo

cosh2 l - 1 - ^ — ^ j - c o s 2 GQ
564 VLF, LF and MF antennas

Design examples showing the agreement between calculated and measured values
of antenna impedance are given in Sections 15.3 and 15.4.

15.2.6 Relation between electrical height and physical height


It is well known from experiment that antenna resonances, i.e. base reactance equal
to zero, occur for physical heights which are less than multiples of X/4. This is
because of 'end-effect', and because tower antennas are not 'thin', the velocity of
propagation down the antenna is less than the velocity of light. The end-effect
which occurs in open-ended transmission lines is due in this case to a decrease in the
inductance per unit length and to an increase in the capacitance per unit length near
the end of the line, and to an increase in the current over that given by a sinusoidal
distribution. In the case of antennas the end-effect produces an apparent length-
ening of the antenna by an amount that depends on the characteristic impedance
of the antenna (the length to diameter ratio) and the current distribution on it.
The experimental model measurements of Brown and Woodward23 showed a
clear dependence on the height diameter ratio (h/d) and that this dependence was
stronger at the second resonance (Ga = 180°) than at the first resonance (Ga =
90°). Their work showed very clearly the influence of terminal effects, which is
especially marked for antennas of large diameter, and these authors tried to esti-
mate the base capacitance due to the metal-capped tube that formed the end of
the antenna. They showed graphs in which this effect had been taken into account.
Various authors (cf. Smith,85*86 Williams,105 Laport56 and Jordan and Balmain,46)
have discussed these effects, and have tried to calculate the impedance using dif-
ferent versions of the transmission-line method for various antennas, in particular
for the full-sized broadcast tower measured by Morrison and Smith.64 This antenna
had a square cross-section 6\ ft on a side, and it was 440 ft high. The tower was
tapered to a much smaller diameter over the bottom 22 ft (5%) of its height, where
it rested on the base insulator. The base capacitance was small but known (30 pF).
King48 has also calculated the admittances of electrically thick cylindrical
monopoles with open and closed ends, and Richmond75 has recently calculated,
using the moment methods of analysis, the admittances of such structures. Clearly,
by computer modelling, the computational tools are available for calculating the
resonant frequencies of monopoles of any thickness or shape and so, theoretically,
one could derive the relation between electrical length and physical length and
diameter.
While such thick structures are of academic interest, they are not generally used
for base-insulated tower radiators. The type of tower measured by Morrison and
Smith is more representative. Accordingly, the author made measurements of the
first and second resonant frequencies for uniform solid cylindrical monopoles
tapered to a point over the bottom 5% of their length for antenna diameters of
about 1°, 3° and 10°. This minimised the base capacitance effect. Also, no base
insulator was employed. The modelled guyed antennas (using insulating guys)
rested on the input measurement terminal of the impedance meter, which was
located directly beneath a large circular-disc ground plane. The results of these
VLF, LF and MF antennas 565

measurements, labelled data points 2, are given in Fig. 15.5. The other selected
data are those of Morrison and Smith (data points 1), and measurements made by
Feldman (data points 3) for second resonance (given by Schelkunoff,81).

0.5 1.0 2.0


ANTENNA DIAMETER (D) DEGREES

Fig. 15.5 Empirically determined relation between the electrical height of a monopole,
graphed as percentage increase over its physical height, H, as a function of its
electrical diameter, D, and height, G.

The curves for G = 90°, 180° and 270° give the percentage increase of the
electrical height G over the physical height H as a function of the diameter D
measured in electrical degrees. The data for 90° and 270° show little scatter,
whereas, even performing the measurements with great care, the data for 180° have
considerably more scatter.

15.2.7 Mutual impedance between monopole antennas


When two or more antennas are used in a directional array, the driving point
impedance of each antenna depends upon the self-impedance of that antenna,
upon the mutual impedance between the antenna and each of the others, and,
in addition, on the complex currents flowing in the various elements of the array.
This results from the natural consequence that an antenna excited so that it has
current flowing on it will induce a voltage on any other antenna in the near vicinity.
This action is as though the two antennas were coupled, and the behaviour that
566 VLF, LF and MF antennas

results can be expressed in terms of the mutual impedances Z 12 = Z 21 existing


between antennas 1 and 2. This is illustrated diagramatically in Fig. 15.6a which
can be represented as far as the voltages and currents which are of concern by the
four-terminal network of Fig. 15.6&.

Fig. 15.6 Mutual coupling between monopole antennas: (a) Antenna configuration; (b)
Equivalent circuit

Zn is the impedance measured at terminals 1 with terminals 2 open; similarly,


Z 22 is the impedance at terminals 2 with 1 open. Z 12 = Z 2 1 is the mutual impe-
dance between the antennas defined by:

Z 21 = - ^ (15.18)
I

where F 21 is the open-circuit voltage induced in terminals 2 of antenna 2 owing to


current It flowing (at the base) in antenna 1. Similarly, Vn is the open-circuit
voltage across the terminals of antenna 1 owing to current I2, flowing in antenna 2.
The mesh equations for Fig. 15.6 are:
Vl =
_IiZ"+hZ» (15.19)

v2 — Az 2 i + / 2 Z 2 2
If we let r = lx\l2 (following Jordan and Balmain46), where in general r is a com-
plex number, then:

7s- = Zn +-Zn (15.20)

j - = rZ2l+Z22 (15.21)

It is clear that the operating input impedances, Vx\lx and F 2 // 2 , are dependent on
the current ratio r, as well as on the mutual impedance which is a function of
physical parameters of the antenna system.
The mutual impedance can be measured by means of an RF impedance bridge,
by measuring (a) the base impedance of antenna 1 with antenna 2 open-circuited;
and (b) the base impedance of antenna 1 with antenna 2 short-circuited.
The first measurement gives the mesh impedance of antenna 1, since, if/ 2 =
0, in eqn. 15.19, the driving-point impedance is
VLF,LF and MFantennas 567

J~ = Zn (15.22)

Under the conditions for the second measurement


Vx =IlZll+I2Zl2
(15.23)
0 =/iZ12+/2Z22
Simultaneous solution of these equations to obtain the driving point impedance
VJIi gives:

r ~ zn -y~ (15.24)
I
I ^22

Thus a measurement of the driving-point impedance VX\IX, together with a know-


ledge of the mesh impedances Z n and Z 22 , gives the information needed to cal-
culate Z 12 .
Under most conditions, the impedances Z n and Z 22 are approximately equal to
self-impedance Zb, except when the antennas are very close together, or very near
a resonant length for the open-circuit configuration (that is, G = 180°).
The mutual impedance between pairs of antennas in a multi-tower array can be
similarly measured by open-circuiting all antennas, and measuring the change in
impedance of a particular antenna, as the terminals of an antenna whose mutual
impedance to it is to be measured is open- and short-circuited.
The mutual impedance between two antennas can be calculated by assuming
a convenient current flowing in antenna 1, with antenna 2 open-circuited, and
calculating the electric field at the position of antenna 2. This calculation gives
the voltage induced in each different length of antenna 2. The voltage thus induced
in antenna 2 has the same effect, insofar as current in antenna 1 is concerned, as
some other induced voltage acting at the reference point (the base of antenna 2)
and this voltage V2 is obtained by integration of these equivalent voltages referred
to this point (cf. discussion by Carter28 and Brown,21).
The results in Fig. 15.7, from Brown (loc. tit), give the magnitude and phase of
the mutual impedance between two electrically 'thin' identical vertical grounded
antennas. The designations on the curves are the electrical heights of the antenna.
Cox29 and Smith85 have given equations for calculating the mutual impedance
between antennas of any height. The curves in Fig. 15.8 give values for mutual
impedance (Xm and Rm) for very small separation between antennas, as required
for design of VLF/LF multiple-tuned antennas.
While these curves are strictly correct only for thin antennas, they can be used
to give values in reasonable agreement with experiment if the electrical height G
is estimated from the physical height H and the electrical diameter/) using the curves
in Fig. 15.5. The error which results if account is not taken of this difference is
shown in Fig. 15.9. The 'theoretical' values shown in the Figure are incorrect
because H was assumed to be equal to G. While 'corrected' theoretical values are
not shown, the difference between the two curves is clearly in the right sense, since
568 VLF, LF and MF antennas

40 80 120 160 200 240 280 320 360


SPACING BETWEEN ELEMENTS - DEGREES

SPACING
300° 340° 380° 420° 460°

-240°
-260°
-280°
-300°
-320°
-340°
-360°

-260
40 80 120 160 200 240 280 320 360
SPACING BETWEEN ELEMENTS "DEGREES

Fig. 15.7 Mutual impedance in magnitude (\Zm\) and phase (&m) for antennas of equal
electrical height G, as a function of spacing between them (degrees)
VLF, LF and MF antennas 569

for H between 90° to 180° the difference between H and G increases as G increases.
The experimental values in Fig. 15.9 were tabulated by Wright (private communi-
cation) from data measured at many sites, data from broadcast station proof of
performance records on file with the Canadian Department of Communications.

600
2 5 10 20 50
SPACING BETWEEN ELEMENTS - DEGREES

Fig. 15.8 Mutual impedance between closely spaced monopoles of equal height, mutual
reactance Xm and res/stance (graphed as a ratio Rm/Rr)

A final check on the calculation of the base operating impedances of the towers,
in a multi-tower array, is obtained during the adjustment of the antenna system and
measurement of its pattern, by use of an in-line operating bridge. This permits
impedance measurement with the towers fed in the desired manner to provide the
required pattern.

15.2.8 Effect of ground conductivity and ground systems on antenna terminal


impedance
The principles of grounding monopole antennas differ according to the frequency
employed and the current distribution on the radiator. At MF, antennas having
heights of X/2 are practical, and since the radiation resistance is high, there is almost
no current flow in the feed to the base of the antenna. Furthermore, at MF, radial-
wire ground systems of the order of one-half wavelength long are practical and
economical. In such systems, most of the electrical flux that causes return ground
conduction currents to enter the base of the antenna as current is collected over the
top of the ground, so current density in the soil beneath is small. At VLF and LF
where the radial wires are electrically short in practice, a considerable part of the
total base current comes from return currents flowing in the ground beyond the
570 VLF, LF and MF antennas
limits of the radial ground system. Also, the return currents flow back to the
antenna at considerable depth under the ground system, and it is therefore impor-
tant to minimise current densities in the soil to reduce losses. The ground wires may
be brought out of and above the surface of the soil at some distance from the base.
This requires the deep currents to rise to the surface uniformly at a distance suf-
ficient to prevent excessive concentration of currents, and hence loss. Also the
radials that emerge and come to the base above the ground form an electrostatic
screen to shield the ground from the intense electric field near the base of the
antenna. This prevents large dielectric losses in the soil.

\\J\J— J
/

tzzj
/

(1=150^—• H /

•/
t/> /

rp ID

-y?f
I r e CV oronirK I^A 1

o j|-)cr»ocTirAi

3
2 y

1.0-
50 60 70 80 90 100 110 120
PHYSICAL HEIGHT (H) DEGREES

Fig. 15.9 Measured values of mutual impedance \Zm\, as a function of the physical height, H,
of radiators of equal height at a fixed spacing (150°). The graph shows the error
that results if the theory neglects the thickness of the antenna

Since the losses which occur when the antenna current returns through the
ground system is one of the most important factors that limits radiation efficiency
at VLF and LF, many theoretical and experimental studies have been made to
minimise these losses in practical antenna design. Wait has been one of the most
prolific authors of theoretical papers on the subject; however, Abbott1 and
VLF, LF and MF antennas 571
63 100
Monteath had studied the subject before Wait and Pope carried out their
analysis. Monteath used the electromagnetic compensation theorem, a method
which has since been used by Wait, Anderson, King and others (cf. Teng and
King,93) to study surface fields and radiation patterns of vertical electric dipoles
over radial ground systems laid on finitely conducting ground. A detailed summary
of Wait's work is given by Watt.102 The NEC computer code referenced earlier (see
Burke et a/.,25'26) has incorporated modelling of antennas near the ground. It is a
rigorously correct mathematical model for structures that are entirely above ground,
but very close to ground. For wires that connect to the ground, the NEC code
assumes that the derivative of current along the wire is zero at the connection
point, to obtain a boundary condition necessary to proceed. This assumption is
rigorously correct only for perfect ground, but it is probably reasonable for mono-
poles that are well grounded. The problem is currently under investigation.

y
5
*
S /
X
1 ~—

A
/
1 JIM CREEK
2 BALBOA \ VL F
J ANTENINA<
/ 3
4
RUGBY
250" UMBRELLA )\ \
C AW A

Y
5 500' TRIANGULAR
6 300' T 1AL IF/\X
7 140' SPIDERWEB

10 100 1000
FREQUENCY kHz

Fig. 15.10 Typical values o f ground loss resistance Rg for electrically short antennas, measured
as a function of frequency for various LF and VLF antenna installations

One of the most detailed experimental studies of ground systems as a factor of


antenna efficiency at MF is the work of Brown et al?2 This work has been subse-
quently referenced by many authors.
The curves in Figs. 15.10 to 15.12 summarise our knowledge about practical
values of effective ground loss resistance Rg referenced to the base of the antenna.
572 VLF, LF and MF antennas

The various curves in Fig. 15.10 show measured values for Rg as a function of fre-
quency for electrically short antennas. These data are considered to be typical of
good installations. The ground loss resistance can be less if the ground screen is
very extensive, or is an ocean; it can easily be much higher for poor installations.
It is seen that Rg varies approximately linearly with frequency.

0.01
GROUND SYSTEM RESISTANCE, R g , OHMS

Fig. 15.11 Ground loss resistance Rg as a function of the total length of wire in the ground
system and frequency for several major VLF stations. The dashed curves are
calculated values

Fig. 15.11 shows measured and calculated values of Rg for some of the major
VLF stations in the world. It is clear that the ground loss resistance varies inversely
with the total length of wire in the ground system, and that hundreds of kilometres
of wire are needed to achieve values of Rg less than 0.1 £1.
The curve in Fig. 15.12 shows values of Rg at MF plotted against the number of
radial wires n in the ground system. In this presentation, Rg varies approximately as
l/n. While Rg is not critically dependent on the length of the radial wires, their
length was between 0.25 and 0.5 wavelength. Values for Rg at MF of 1-2 SI,
referenced to the current loop, are typically quoted in the literature as being
representative of good MF broadcast antenna systems.
Obviously, the length and number of radial wires should be as great as possible,
but practically more than 120 radial wires are seldom used, and for single tower
radiators the length of the radial wire should be at least 2-4 times the height of the
VLF, LF and MF antennas 573

radiator. The VLF antenna systems with very extensive top-loading, the radial wire
system should extend a tower height or so beyond the outer limit of the top-
loading. At MF where two or more towers are used to form a directional array,
each radiator must have its own radial ground system and adjoining systems should
not overlap but are terminated in a common bus. A grid wire system might be a
better arrangement in this case, but since grid wire systems are not used, there are
no examples for comparison.

/
> BELROSE AT 2 8 M H T
/
BROWN ET Al AT ; JMHz

/
/
in
4

y
c

y
y

500 250 100 60 40 20 10 8


NUMBER OF RADIALS (n)

Fig. 15.12 Ground loss resistance Rg as a function of number of radial wires in the ground
system over sandy soil (frequency about 3 MHz)

A recent study (Tippe,96) has shown that for radiator systems between 3X/8 and
5X/8, the conventional approach at MF of using 120 radials between 0.25X and
0.4X long may be over-dimensioned, and as few as 50 radials 0.2X long may be just
as effective. Monteath106 has also shown that the efficiency of an earth system is
not greatly reduced if only 72 radial wires are used.
Wire size apparently has a negligible effect on the effectiveness of the ground
system and, except for very high power stations, the wire size is chosen for
mechanical strength. Wire of size AWG no. 10 or larger is typically employed. The
wire is usually buried 10-15 cm, or much deeper to permit cultivation of the soil.
When such deep burial is required, the wire should descend to its required depth
on a smooth gentle incline from the tower base (Head40). Usually, bare solid copper
is employed, but insulated wire has also been used; this increases its mechanical
574 VLF, LF and MF antennas

strength and protects it from corrosion. In fact, insulated wire is claimed to provide
better stability with seasonal changes of earth conductivity (Sather,77). Alterna-
tively, a counterpoise is sometimes employed, at least over some radius about the
towers in a multi-tower array, for the pattern stability.

15.2.9 Effect of the finite conductivity of the earth on the vertical antenna
radiation pattern
The finite conductivity of the earth not only affects the input terminal impedance
of antennas, but the ground in front of the antenna has a major effect on the sky-
wave field strength launched at a particular angle of radiation above the horizon.
The vertical pattern shapes, which were sketched in Fig. 15.2, have been calculated
assuming a perfectly conducting flat earth.
In Section 15.2.2, we noted that the unattenuated field strength at 1 km in
front of the antenna for 1 kW of transmitter power was a figure of merit for a
particular antenna system, since this value could be compared with that expected
for an antenna with no losses. This field strength is easily measured, and it is
the field strength that would be used for purposes of predicting ground-wave
propagation. The radial ground system and the ground surrounding it, within a
radial distance of a wavelength or so, are effective in determining the terminal loss
resistance. This is not the situation for launching of skywaves. The reduction
in the skywave field strength because of the imperfectly conducting curved earth
is influenced by the ground well beyond the ground screen; in fact, for radiation
angles less than 5° above the horizon, the ground conductivity fifty or more wave-
lengths in front of the antenna is effective in determining the vertical radiation
pattern.
Thus, while the terminal-loss (ground-loss) resistance limits the total power
radiated, and hence the field strength at a particular radiation angle, the shape of
the vertical pattern is determined by the ground in front of the antenna far beyond
the limits of the antenna and its radial ground system.
While qualitatively the effect of the finitely conducting curved earth on the
vertical pattern of antennas is well known, and has been discussed by several
authors, quantitatively, the theoretical calculation of these patterns is not straight-
forward. Furthermore, the ability of an antenna to launch skywaves at low
elevation angles cannot be measured directly since, by employing a receiver in
an aircraft, the signal received is the total field comprising ground and skywaves.
While the vertical pattern for launch of skywaves can be rigourously calculated
employing a Fresnel reflection-coefficient model for the ground, or the SommerfeId-
Norton integrals (cf. Trueman and Kubina97), for propagation over a flat earth,
there is a discontinuity in the skywave field strength at grazing incidence (see
Fig. 15.13). That is, the skywave field decreases to zero as the radiation angle \p
approaches zero (angle of incidence 6 approaches 90°). Note that the field for
perfect ground does not decrease to zero, but this is only because the ground wave
was included in the calculation of the far field. For a perfect earth, the effect would
monopole r field
point

ZS Televation

ground screen
of 8 radials
I I i 1 1 I

..x-.-x— -x perfect ground


0.5 0.5 - -
perfect ground ^ s ^ * . /

! Fresnel Method X
en
CD 0.4 1 * I CD 0.4 S-N Method _
S- -- .—. Space wave pj U 8
S-
4-* / i^S-N Methoa +->
CO ground wave
(b) -a
0.3 0.3
X

\ (0
X
\
0.2 0.2 - \ _

:ive F- i e l d
CO S-N Method
CD
CD
x / s p a c e wave only
O.I O.I -

AJk 1 1 1 i i j
10 30 50 70 90 0 8 5 86 87 88 89 9 0 i
e degrees e degrees
Fig. 15.13 Calculated elevation pattern for a vertical monopole (H =201°) with eight radials at 28X distance for a ground conductivity of 20
mS/m at a frequency of 860 kHz; (a) Monopole with co-ordinate system; (b) Elevation pattern; (c) Detail of elevation pattern near
grazing incidence
576 VLF, LF and MF antennas
be the same, except that it would be more closely confined to angles near to grazing
incidence. In the real world situation, it is clear that the skywave field is not zero
when \jj = 0.

f=20kc/s
O.l- _PAST0RAL LAND.
< €=I5€ O

cr=5xl6 3 S/m
a = 4 / 3 x 6 3 6 0 km
V200

0.01-
15 10 5 0 -5 -10 -15 -20
ANGLES ABOVE HORIZON, * , (DEC)—•*

LJ

0.01
10 5 0 - 5 -10 -15 -20
ANGLES ABOVE HORIZON,»//, (DEG.) —

Fig. 15.14 Antenna factor F, as a function of frequency and radiation angle \p for: (a)
Pastoral land (o = 5XW-3 S/m); (b) Sandy soil (o = 1 X W'3 S/m)

The real earth is not only finitely conducting but it is curved, and account needs
to be taken of both these factors, especially for frequencies below MF and for
estimating radiation patterns at low angles above the horizon. Wait and Conda101
have considered skywaves incident on a finitely conducting earth; and evoking
reciprocity, their numerical analysis, which includes also atmospheric refraction,
VLF, LF and MF antennas 577

can be used to deduce the effect that the curved earth has on the effective antenna
radiation pattern. The curves in Fig. 15.14, which were deduced by Belrose11 from
their work (see also CCIR Report 265-4, 1978), can be used to take account of the
effect on the antenna pattern of a finitely conducting curved earth. Results are
shown for two ground conductivities, pastoral land (5 x 10~3 S/m) and sandy soil
(1 x 10~3 S/m). The factor F, when multiplied by the field strength that has been
calculated for a perfectly conducting flat earth at the corresponding radiation
angle, gives the expected field strength that would be radiated at that angle. Knight
(see CCIR Report 401-3, 1978) has made similar calculations for LF/MF broadcast
frequencies (0.2-1.5 MHz). The negative angles of radiation apply when waves
diffract around the curvature of the earth, and which are defined as the angular
distance between the receiving antenna and the tangent point for the downcoming
skywave that has been reflected by the ionosphere.
It is clear that the modification to the vertical radiation pattern is that of a
reduction in field strength, and this reduction is greatest for radiation angles near
grazing incidence and for land of poor conductivity. The reduction is greater at
MF than at VLF.

15.2.10 Conductor and tuning-coil losses


Because of the rather large currents on VLF and LF transmitting systems, the loss
resistance of the tuning inductor necessary to resonant electrically short antennas
can be an important consideration. Furthermore, because of the relatively large
size of antenna tuning inductors, and the high voltages that result from the flow of
high currents, the construction of these inductors and of the helix house that
shields them can present special problems. This subject will be considered briefly
in Section 15.9 but, for the present, we are concerned with typical losses, since
when we consider representative antenna designs in the next Sections, it will be
necessary to estimate these tuning inductor losses.
The helix inductors used for VLF and LF antenna systems are constructed
from litz wire cables, which are wound with multistrand groups over an insulating
core. Litzendraht wires having hundreds to thousands of strands, each of 0.1-0.25
mm diameter, are used for low-loss antenna systems. Coil Q-factors

G = J-T (15.23)
K
c

of 1000-3000 are practically achieved, for inductances up to 10 mH. At VLF,


helix (^-factors of 2000 are representative of good antenna installations, and at
LF about 1000-1500.
Litz cables are not useful for frequencies greater than 500 kHz. Inductors used
at MF are wound from 0.5-2.54 cm-diameter copper tubing, which is sometimes
silverplated. Coil Q-factors of 500 are typical, but inductor losses are small, since
inductances required are not large (100 JUH or less).
Watt102 discusses in more detail coil quality and losses and Martin and Carter61
reference a number of papers that discuss helix inductors and insulator losses.
578 VLF, LF and MF an tennas

15.3 Practical antenna design: Loaded antennas

15.3.1 Types and methods of feed


Vertical radiators which are electrically short have a low radiation resistance and a
relatively high capacitive reactance. At frequencies near the operating frequency,
the antenna can be considered as a lumped circuit, which consists of a resistance
and a capacitance in series. The base resistance Rb includes the radiation resistance
Rr and the effective ground loss resistance Rg, and the reactance Xb is the capaci-
tive reactance at the operating frequency. It is clear that for this antenna to take
power, the capacitive reactance must be tuned out by a suitable inductor, and the
resulting resistive impedance matched to the transmission line feeding the antenna.
This tuning inductor introduces addition loss resistance Rc, and the object of design
is to obtain the highest possible practical ratio of radiation resistance to total loss
resistance Rt.
It is clear that some form of top-loading will increase the capacitance of the
antenna and so reduce the capacitive reactance at the base of the antenna, increase
the current area on it and so increase the radiation resistance, which results in a
marked improvement in radiation efficiency and bandwidth.

7////////////////7/7/777 77 7 7/

(o) (b)

(
(c) (d)

(f)

Fig. 15.15 Method of feed for various grounded and ungrounded antennas of the T type
(a) to (d); and umbrella types (e) and (f)

Fig. 15.15 illustrates several basic types of top-loaded antennas and also dif-
ferent methods of feeding base-insulated and grounded tower radiators. The arrange-
ment in (a) is the one that is most commonly used. It is a base-loaded T-type
radiator; sometimes the inverted-L type is used. The horizontal arms of the T are
VLF. LF and MF antrnn** 579
suspended between two supporting towers. While the sketch illustrates it as a
single-wire antenna, in practice, the top-loading and the vertical radiator can take
the form of a cage of wires to increase their capacitance and reduce the base termi-
nal reactance of the antenna. Sometimes the top-loading is configured into the
form of a delta suspended between three towers; or it can take the form of a
diamond requiring four support towers.
The sketch in (b) illustrates an antenna type that is seldom used because of the
electrical and mechanical difficulty in installing and tuning a coil at the top of a
tower radiator. Since the coil is tuned so that the base reactance is zero, this arrange-
ment yields the greatest possible current area (for Ga < 90°) on the vertical radiator,
for an antenna of a given physical size, and hence a maximum value of the radiation
resistance. The improvement in radiation efficiency is, however, partly compromised
by the fact that a larger tuning inductance must be used compared with the con-
ventional base-loaded radiator.
The sketch in (c) has been said (Ray,72) to be electrically equivalent to (b), but
with the advantage that the tuning coil is at the base of the antenna. Using the
tower as a transmission line, the antenna is effectively fed at the top, between the
tower and the top-hat terminals. While this can result in an increase in the radiation
resistance over that for the simple base-tuned insulated tower, very heavy top-
loading must be employed. In practice, in the experience of the author, the effective
radiation resistance can be less than that for conventional base-loading. The require-
ment for 'heavy' top-loading can be seen, at least qualitatively, since in (c) if the
top-loading is reduced to zero, we no longer have a radiator; whereas in (a) with
no top-loading the antenna becomes a simple straight base-loaded vertical
monopole.
The arrangement in (d) is an alternative method of feeding a grounded tower; in
this case, the entire antenna is a DC ground. This is advantageous for two reasons:
no base insulator is required to support the tower and the entire antenna is grounded
for protection against lightning. While a capacitance match network is illustrated,
on the assumption that the input impedance is inductive, this is only the case for
very short antennas, as we shall see later when we consider the folded-monopole
antenna. The physical arrangement in (d) is to surround the tower by a cage of
wires, employing a conducting outrigger at the top of the tower and insulated
outriggers at suitable intervals along its length. The feed is between the skirted
end of the cage and ground.
The final two sketches illustrate two arrangements for feeding grounded tower
umbrella-type top-loaded radiators. In the umbrella arrangement, the top-loading
takes the form of a number of active guys, attached to the top of the tower and
insulated at their lower ends. These active guys are arranged symmetrically around
the tower, and while only two are shown, usually 4, 8, 12 or 24 active guys are
employed. The antenna type in (e) is a folded-umbrella-type antenna (cf. Nolan66);
in (J) there are no connections made to the supporting tower. A cage of vertical
wires, shielding the tower from currents flowing on the cage, connects the base-
feed terminal to the umbrella top-hat.
580 VLF, LF and MF antennas
153.2 T-type antenna
The cage T-type radiator, a model of a particular installation, shown in Fig. 15.16,
is a particularly simple antenna to analyse, because it is symmetric in shape and

Fig. 15.16 A model of a cage T-type antenna suspended between two self-supporting
grounded towers

form. In the installation shown, the entire system, see Fig. 15.17, is constructed
from four-wire cages of 6.55 mm-diameter wires spaced 1.22 m from each other.
The flat-top is suspended between two 91.4 m grounded self-supporting steel
VLF,LF and MF antennas 581
towers. The average height of the flat-top is about 82.3 m, the length of the radiator
and of each arm of the flat-top is 76.2 m. Measured values (at full scale) of base
reactance Xh and resistance Rb are shown in the Figure. The resistance values for
frequencies less than 100 kHz have been extrapolated. The curve for radiation
resistance Rr has been calculated, but the values have been verified by measure-
ments at model frequencies and at full scale.

-182.9 m

4-WIRE CAGES OF 6.55 mm DIA.


91.4 ml WIRE IN FORM OF SQUARE
1.22 m PER SIDE.
76.2 m

2000

O.I
70 I00 200 500
FREQUENCY kHz

Fig. 15.17 Base resistance Rb, reactance (—jX^), and radiation resistance (Rr) for a cage T
antenna having the dimensions shown, as a function of frequency

While this antenna has already been analysed by Laport,56 it has been studied by
the author at full scale and by experimental modelling; and so, in the following, we
will carry out a detailed analysis of it to illustrate the application of the various
equations given in Section 15.2. The frequency is 100 kHz and the wavelength is
therefore 3000 m. At 100 kHz, the vertical height Hv of the radiator and the length
of each arm of the flat-top Lh are 9.15°. Assuming 5% lengthening to account for
end effects:
582 VLF, LF and MF antennas
Gv = Lh = 9.15(1.05) = 9.6°
The characteristic impedance of the vertical radiator is

Zg = 60

where a is the effective radius of the element. For an effective radius of 0.72 m,

In— 1 1 = 220 ohms

The characteristic impedance of the flat-top can be calculated from the equation
for the impedance of a conductor of effective radius a at an average height h above
the ground:

Zg = 6o(ln^j = 60 ( i n ^ f ) = 326 ohms


The reactance of one horizontal arm of the T at the junction is
Xh = — J'ZQ cot Lh

= —/ 326 cot 9.6 = - / 1927 ohms


With the two arms in parallel, the total reactance X of the flat-top is
X
— = — / 964 ohms
The reactance of the flat-top (—/964) is equivalent to the reactance of a length 6 of
characteristic impedance ZQ given by
964 = Zl cot 6 = 220 cot B
This gives 6 = 12.8°
The base reactance is then
Aj, — 7^0 COT, \Ja

= - ; ( 2 2 0 ) cot (9.6+ 12.8) = - / 5 3 3 ohms


This value should be compared with 530 O measured for the modelled antenna
and 640 12 at full scale.
The current area on the vertical radiator is
180/cos (Ga - Gv) - cos Ga\ A
A = —I degree-amperes
n \ sinGa ]
180/cos (22.4 - 9.6) - cos 22.4\
A = 1= 7.6 degree-amperes
71 \ sin 22.4 /

Rr = 0.01215 A2 = 0.01215 (7.6)2 = 0.7 ohms


VLF, LF and MF antennas 583
If Rg = 1 ohm (see Fig. 15.10), and the tuning-coil £>-factor = 1000, the radiation
efficiency is:

= Rr (100) = 0.7(100) = 3
Rr + Rg+Rc 0.7 + 1.0 + 0.53
The antenna Q-factor is

X> 533
Qa
Rt 2.23
The antenna bandwidth is therefore

Since the antenna system is matched to the output impedance of the transmitter,
the operating bandwidth is
BW (operating) = 2 (0.42) = 0.84 kHz
For a transmitter power of 25 kW

/25,000
#» J-T^T = 106 amperes

The antenna base voltage is


Vb = 4 X 5 = 106(533) = 56,434 volts (56 kV)
To calculate the self-resonance of the antenna system, we proceed as follows. At
the fundamental frequency of the system / = / 0 , Xb = 0 and the top-loading
reactance is

Now this reactance Xo is one-half of the reactance of one arm of the T at the
junction. Hence
.ZpCQtL,, _^ 0
/ - / z 0 tan Crv

Since the horizontal arm and the height of the radiator are each 76.2 m (adding
5% for end effects),

L°h = G?v = 7 6 . 2 ( 1 . 0 5 ) ^ ( 3 6 0 ) = 96/ 0

Substituting this value in the equation above gives


584 VLF, LF and MF antennas

220ta

and, solving for/ 0


/ 0 = 0.424 MHz (424 kHz)
Fig. 15.17 shows that this value for / 0 is approximately that measured (actually
/o measured was 390 kHz). It should be noted that since Lh = Gv, the solution for
/o was particularly simple. In general, this equation would have to be solved by
iteration.

15.3.3 Transmission-line radiator


Consider an L-type antenna of height h and top-loading length /. We have already
noted that if the top-loading is increased to the point where the antenna is resonant,
i.e. its input reactance Xb = 0, this results in increased current area on the radiator.
Furthermore, if h < /, the form of the structure resembles a shorted transmission
line, consisting of the shorting stub (the radiator), a horizontal conductor and its
reflection in the ground, see Fig. 15.18a. In this sketch ZT is an open circuit or a
very small capacitor which can be used to tune the antenna over a narrow band of
frequencies, and 1 + h ~ A/4 when the antenna is resonant. Since the top-loading
conductor contributes very little to the field radiated by the antenna, it can be
formed into a circle, or wound up like a spiral, as in (b) and (d).
A practical arrangement is illustrated in Fig. 15.18c, where the shape of the
top-loading is octagonal, and, considering the antenna as a transmission line, it can
be fed at a distance X from its 'shorted' end to provide a match (50 £1 impedance)
for feeding the antenna by means of a coaxial-cable transmission line.
The ring form of the antenna (b) has been the subject of some controversy in
the literature (Boyer,19 Burton and King,27); however Belrose12 and others have
considered it as one form of a class of transmission-line radiators, and he modelled
the antenna and measured its radiation parameters to verify this interpretation.
The antenna modelled was electrically very short, h/X = 0.007 (H =2.56°), and,
while one of the design parameters was empirically determined, and may not apply
for antennas of greater electrical height, we summarize below this analysis of the
antenna:
The current area A on the vertical section of the antenna is:
180
A = — sin Gv degree-amperes
7T

where Gv = 1.05Hv = 1.05(2.56) - 2.7°

180
A = — sin 2.7 = 2.69 degree-amperes
n
and Rr = 0.01215(2.69)2 = 0.09 ohms
VLF, LF and MF antennas 585
Using the measured radiation efficiency of 1.08% for the modelled antenna, the
formula

r? = ^ (100)

gives Rb, which is equal to the sum of the radiation resistance and the ground
loss, as 0.8 ohms.

ZT h<< X

c) d)

Fig. 15.18 Various configurations for a transmission-line radiator (h < \/4). The sketch at (c)
shows a practical method of matching and feeding the radiator

The characteristic impedance of the transmission line is given by

Z o = 60 In — ohms
a
which for the antenna modelled equalled 293 £1.
The Q-factor for a quarter-wave transmission line is
2nf0Z0
Q =

and therefore the antenna bandwidth is


Re
BW = £ = hertz
Q
where Reff = effective loss resistance per unit length along the transmission line
586 VLF, LF and MF antennas

v = velocity of propagation down the line (approximately equal to


0.975 times the free-space velocity)
For the purposes of this calculation, consider that the total antenna resistance
(Rb — 0.8 12) has been distributed over the length / of the radiator h, hence the loss
resistance/unit length can be calculated, and the equivalent loss/unit length dis-
tributed over the length of the transmission line is
R 1
jg^ = — __ ohms/m

For the antenna modelled Reff = 0.3 O/m, and the antenna BW(calc) at the model
frequency of 7.785 MHz is 48 kHz. The measured BW was 43 kHz.
The impedance of the open end of the transmission line is

2ZO2
Rs =
^~T ohms
This is an important parameter in antenna design, since the voltage at the open end
of the line can be estimated by assuming that the transmitter power is dissipated in
this resistance.
The effect of tapping along a line a distance $X (electrical degrees) is equivalent
to transforming the antenna impedance Rh to the line impedance Rx, and this
impedance step up can be calculated from the transmission-line equation

Rb sin2 (pX)eff
where (pX)eff is used in the calculation instead of the actual electrical distance
PX. This is because, for small PX, (fiX)eff is greater than PX due to coupling between
the driven and shorted (radiating) elements of the transmission line. The relation
between PX and (pX)eff is given in Fig. 15.19 for the modelled antenna. It should
be noted that for PX greater than 5°, (pX)eff is equal to PX.
The curves in Fig. 15.20 give calculated values of the radiation efficiency for a
low-loss (the one modelled), and for a very low-loss (essentially including con-
ductor losses only) transmission-line radiator versus the height of the radiator
(in electrical degrees).

15 3.4 Diamond antenna


Laport56 has given reactance versus frequency values for a diamond antenna sup-
ported by four 400 ft (122 m) grounded steel towers. Fig. 15.21 shows dimensions
and impedance versus frequency calculations based on these data for tower heights
of 1000 ft (305 m). Naturally, we could have scaled the data (up or down) to any
other physical size, or we could have drawn curves for Rr and Xb versus hj\ instead
of frequency. In this case, we could not have shown a curve for Rb since Rg is a
function of frequency (see Fig. 15.10).
VLF, LF and MF antennas 587
15.3.5 Large VLF antenna
Fig. 15.22 shows a pictorial sketch and impedance frequency data for one of the
two largest antennas in the world. This one is the US Navy transmitter at Northwest
Cape, Australia (NWC). The top hat comprises six diamond antennas suspended
from twelve 303 m towers arranged about a central tower of height 390 m. All
towers are grounded. The down-leads are arranged symmetrically about this central
tower (48 wires around this tower). The effective height of the antenna, even with
this massive top-loading, is only one half the average height of the 13 towers due to
practical considerations such as sag in the top-hat, and due to the effect of grounded
towers and their extensive guy wire systems. Each tower has three sets of guys at
each of four or five levels (see Wheeler,104).

1
DEGREE

«•- \
\
\
=8. \

in \

0.3 0.5 I 2 5 10
fix ( EXPERIMENTAL ) - DEGREES

Fig. 15.19 Empirical relation between fix' (the distance measured from the shorted end in
degrees) to pxefp the value used to calculate the impedance at the feed point,
for a transmission-line radiator of type shown in Fig. 15.18c

At an operating frequency of 15 kHz,


Xh = - / 5 7 f t

Rr = 0.14 ft
Rb = 0.165 ft
If the helix tuning coil factor is 1500, we can calculate the following:
Radiation efficiency = 61%
Antenna g-factor Qa = 273
588 VLF, LF and MF antennas
Antenna bandwidth BW = 55 Hz (Operating BW = 110 Hz)
Base current (for 2 MW input) = 3098 A
Base voltage = 177 kV
Particularly impressive are the reactive power of 624 MVA in air dielectric, and the
real power of 2 MVA delivered to a resistance of 0.21 £2.

20
j

10
/
/ f

/*>
y
A
Yy
AV
LI A
Fig. 15^20
1 I 50
RADIATION EFFICIENCY (%)
100

Graph showing how the radiation efficiency of a transmission-line antenna would


be expected to vary with the height above the ground plane for a low-loss and
very low loss antenna

15.3.6 Umbrella antenna


The top-loading for an umbrella antenna, illustrated in Fig. 15.23, consists of a
number of wires strung obliquely to ground from the top of the tower, and insulated
from ground. The important parameters, see Fig. 15.24a, are the height of the
tower h, the horizontal distance d from the base of the tower to the extremities
of the guys supporting the umbrella wires, and the vertical distances measured
from the top of the tower to the height at which the umbrella wires are broken
by an insulator. This antenna was first used by Smith and Johnson88 at broadcast
frequencies, it was investigated experimentally by Belrose et al.1* and by Gangi
VLF, LF and MF antennas 589
3S 84
et al., these authors, along with Smeby examined the antenna theoretically.
Smith and Graf87 have experimentally investigated umbrella antennas using multi-
wire rib-construction to increase the top-hat capacitance but the same effect can be
achieved by using more umbrella wires.

1000'

2000 •

1
i
I
KDO'
8" 4 -1000' TOWERS
(GROUNDED)
jic
i
* 2250' •j
°000

1000- \ 1
10
—Xr —t—'
X j. J
g 500- \ -5

°X \ R
b//
A
7 200- V /Rr
-2
LJ
O / /\ \
2,oo- ' V-
^—^— \
1 V
% 50 0.5

/ /

['
}

/
20- -0.2

/
10- 1 (__ -O.I
20 50
FREQUENCY kHz

Fig. 15.21 Base res/stance R^, reactance —JXb, and radiation resistance Rr for a diamond
antenna having the dimensions shown, as a function of frequency

Figs. 15.24a, b and c illustrate by the arrows the phasing of currents on umbrella-,
T- and I^type top-loaded antennas. It is clear that in (b) and (c) the currents on
the top-hat do not interfere with currents on the radiator, but this is not the case
for the umbrella antenna. In fact, if the number of umbrella wires is sufficient,
and/or if these wires are connected by a skirt wire which is equivalent to increasing
their number, the top-part of the radiator, under the hat, can be effectively screened,
and radiation is due only to currents flowing on the lower part of the tower (having
a height h — s).
590 VLF, LF and MF antennas

Thus the radiation resistance increases as the length of top-loading increases, but
only to a point, beyond which it decreases. The height for i? r (max) is s/h — 0.43.
Since the antenna reactance decreases continuously with increases in top-loading,
the maximum radiation efficiency can occur for top-loading somewhat longer than
that for the maximum radiation resistance.

15 20 25
FREQUENCY kHz

Fig. 15.22 Base resistance Rb, reactance —jX^ and radiation resistance Rr for the USN
VLF station at NWC, Australia having the dimensions shown, as a function
of frequency (after Patch, private communication)

While there are many parameters that can be varied, in practical design, d should
be as large as is structurally feasible. The design data to follow have been measured
by antenna modelling for djh = 1.4, which is considered to be a practical design.
The experimental modelling measured not only antenna resistance and reactance
for various top-loading arrangements, but transmitter power and field strength
were also measured, with reference to a quarter-wave monopole,and so the radiation
and ground loss resistances were deduced.
VLF, LF and MFantennas 591
Fig. 15.25 shows the amount of top-loading s/h0 needed to resonate an antenna
of height h0 at the operating frequency / 0 . The parameter n is the number of
umbrella wires. The curves show the relation between s/h0 and ho/Xo. For a given
antenna configuration, and s/h0 chosen, the curves in Fig. 15.26 show the mag-
nitude of the radiation resistance Rr at frequency/ 0 . For short antennas shorter
than hQ

where h = height of the antenna and h0 is the height required for resonance.

Fig. 15.23 Model of an umbrella-type antenna with 24 skirted radials (see Fig. 15.29 for
full-scale dimensions)

Naturally, if h is less than h0, the antenna reactance Xa is not zero. The curves
in Fig. 15.27 give values of the antenna reactance Xa for various values of h/hQ
(or /// 0 since ratios of frequencies and heights scale in proportion).
If the antenna is reactive, and if Xa is very much greater than Ra, then the
antenna Q-factor and bandwidth can be calculated as discussed above. However, if
the antenna is operated very near or on its self-resonant frequency, then the antenna
Q-factor cannot be simply calculated. Fig. 15.28 shows how the antenna ^-factor
depends on the parameter s/h0. The (2-factors shown are those for the modelled
antenna where Rg = 0.75 O; hence for an antenna at full scale having a different
loss resistance,

R a (model) \
Qa = Qmo,del
R a (full scale) J

Obviously for s/h0 = 0, the antenna must be quarter-wave resonant and its
radiation resistance Rr = 36 O (see Fig. 15.26), and its Q-factor will be low (see
Fig. 15.28). For s/ho>Q, the antenna now becomes electrically short and Rr
592 VLF, LF and MF antennas

//////////////////////////////////////y

(a) (b)

7777777/7777/7////
(O (d)

Fig. 15.24 Sketches illustrating phasing of currents on umbrella, T- and L-type radiators.
The sketch in (a) illustrates base insulated series feed; that in (d) a grounded
tower folded umbrella antenna

0.7-

0 6
\ \ \
\l6
V
\
1 \ \\ -—-
\\ \ \
2
3T SK RT ED
- s KIR TEC)
0 5

1 >

0 2—

0 I
1N \\
\

k
\
\

Fig. 15.25 Empirical relation between the amount of top-loading (measured by the parame-
ter s/h0 and number of radial wires (n)) required to resonate an umbrella antenna
of height ho/\
VLF, LF and MF antennas 593
decreases and the antenna Q-factor increases. Beyond a certain point it will no
longer be possible to resonate the antenna at the operating frequency and its input
impedance will be reactive as well as resistive.

3 5 7 10
RADIATION RESISTANCE Rr

Fig. 15.26 Empirical relation between the radiation resistance Rr and the amount of top-
loading (measured by the parameter s/hQ and the number of radial wires n)
required to resonate an umbrella antenna

The data given in Fig. 15.29 show measured values for base reactance, base
resistance and radiation resistance for a full-size umbrella antenna having the
dimensions shown. Since the antenna is short, the amount of top-loading has been
increased beyond that for maximum Rr.
To illustrate the use of the graphs suppose we design an antenna for a frequency
of 200 kHz:

/ 0 = 200 kHz
Xo = 1500 m
for s/h0 = 0.43, from Fig. 15.25
hlh0 = 0.1
or h0 = 0.1(1500) = 150 m
594 VLF, LF and MF antennas

According to Fig. 15.26 the radiation resistance is 6.75 £2. If the ground loss
resistance is 2.5 O, the radiation efficiency is
Rr(l00) 6.75(100)
= 73%
Rr+Rg+Rc 6.75 + 2.5
Note that Rc, which is equal to the coil tuning loss, is zero since the antenna is
resonant.

Fig. 15.27 Empirical relation between the base reactance —JXa and the height of the antenna
h/h0 and amount of top loading

The antenna g-factor is given in Fig. 15.28. The Q for the model was equal to
20, but since Rg for the model and Rg for the antenna under consideration are
different

and the antenna bandwidth is

BW = =
77^ 12.3 kHz (operating bandwidth = 24.7 kHz)
lb.2

To illustrate the use of the curves for h<h0, suppose that h = 83 m, then
VLF, LF and MF antennas 595
h 83
= 0.055
1500
for eight umbrella wires, and if s/h0 = 0.43, as before, then ho/Xo = 0.1.
Therefore

/
and according to Fig. 15.27
Xa = --/215 ft

'/
07
/ /

/

/
06
Cy

//
1v
04

0 3~'

o ?—
/
/
V
/ / /

n—
15 20 30 40 50 70
ANTENNA Q-FACTOR
Fig. 15.28 Empirical relation between the antenna Q-factor for a resonant umbrella top-
loaded antenna, and the amount of top loading (measured by s/h0 and number
of radial wires n)

The radiation resistance is

Rr = ^] = 6.75(0.55)2 = 2 12

Naturally, we could design the antenna to reduce this reactance by increasing the
number of umbrella wires or their length.
The largest umbrella antenna in the world is that used for OMEGA transmissions
at Tsushima, Japan. The antenna employs a single insulated tubular mast, 450 m
596 VLF, LF and MF antennas

high, with a top-hat consisting of sixteen 4 cm-diameter conductors radiating


outwards and downwards from the top of the mast. The active length of each top
conductor is 315 m. The earth system consists of 90 radial conductors, each 200 m
long, spaced at 4° intervals. The parameter d/h is equal to 0.45; this is much less
than that modelled by the author (in fact, d should be greater than h). At 10.2 kHz,
the measured parameters are: Xa = —/488 $l\Rr = 0.084 12, radiation efficiency 17
= 12%; Pt = 88 kW for a radiated power Pr = 10 kW; antenna bandwidth = 15 Hz;
antenna base current 345 A for a base voltage of 160 kV (Swanson et al91).

1
= I78' = 25O'

U—d = 35O'—-I

50 100
FREQUENCY kHz

Fig. 15.29 Base res/stance /?&, reactance —jXb and radiation resistance Rr as a function of
frequency for an umbrella antenna having the dimensions shown

Folded umbrella antennas have been used at broadcast frequencies, where


h/X0 =0.1 (20 m at 1500 kHz), which can be designed to match the impedance
of the feeder transmission line, requiring no base insulator and no antenna matching
(see Fig. 1524d and Nolan66).
VLF, LF and MF antennas 597
15.3.7 Multiple-tuned VLF antennas
Multiple tuning was first suggested by Alexanderson3 for the purpose of reducing
ground loss of VLF antennas of the long inverted L-type. The idea was to employ
a number of tuned down leads, each with a substantially independent ground
screen, but all sharing a common top-loading. If the individual antenna systems are
small, all down-leads will carry equal in-phase currents. Martin and Carter61 show
a number of VLF transmitting antennas of the type described by Alexanderson, as
well as a multiple-tuned umbrella type. Except for these early VLF antenna instal-
lations, multiple tuning has not been used to any extent in recent years, although
the principles of the idea have not been forgotten.
On the assumption that the down-leads carry equal currents, and that the
current supplied by the transmitter is I/N where TV = number of down-leads,
Laport56 and Martin and Carter61 have argued that the new radiation resistance
Rrr is related to the single tuning Rr by the equation Rrr = RjJV2. However, since
the top-loading is shared by several down-leads, the base reactance required for
each is higher, and a larger coil must be used. For a given coil Q-iactor, the coil
loss resistance in each down-lead will be higher, but the current in the down-lead
is less. It is further argued that if the ground loss resistance is Rg for single-tuning,
then it is Rg/N for multiple tuning. The total antenna resistance with multiple
tuning Rtt is then, in general, less than RJN2, and hence RrrlRtt>Rr/Rt and so
the radiation efficiency is improved.
Laport56 states that the full explanation is more complex than indicated above,
where only the basic principles are explained. Several factors in fact limit the
extent to which multiple-tuning can be used to advantage. Several down-leads subtract
somewhat from the effective height established by the flat-top. As the number of
down-leads increases, their currents tend to overlap to a greater degree, with the
result that the ground loss resistance is not reduced by the factor indicated in the
simple analysis. Also, multiple-tuned antennas are difficult to adjust when a change
in operating frequency is made, and the use of multiple helix coils and helix houses
increases the cost of the antenna.
A more recent analysis of multiple tuning by Ray,72 using the mutual impedances
between the multiple elements (see Fig. 15.8) instead of the simplified analysis
based on proportion of antenna currents, has concluded that the radiation resis-
tance with multiple tuning increases almost proportionally with the number of
elements used. This contradicts the deduction above that the radiation resistance
increases as the square of the number of elements.
Ray72 considers that multiple tuning can be used to advantage for very short
antennas, since the resistance Rtt at the feed point is higher, and the antenna
bandwidth is increased. The principle of multiple tuning has been applied in the
design of a short LF broadcasting antenna, where high efficiency and a bandwidth
required for broadcasting were desired (see Section 15.5.2).
A type of antenna patterned after Ray72 is illustrated in Fig. 15.30. This antenna
not only uses multiple tuning to increase the base resistance of the feed point, but
also an arrangement that employs a loading coil which is effectively in series at the
598 VLF, LF and MF antennas

top of the tower, but physically placed at ground level. This antenna element,
as described earlier, is effectively fed at the top between the tower and the top-
hat terminals. It is argued that this results in the maximum radiation resistance for
a radiator of a given height. Ma and Fitzgerrell31 have used an antenna of this type
on the roof of a van, and while the application was at HF, the results of their
measurements give further insight into this antenna.

Fig. 15.30 Method of feeding a multiple-tuned diamond antenna supported by four grounded
towers (after Ray,72 1972)

15.4 Practical antenna design: Unloaded antennas

15.4.1 Vertical monopoles h < X/4


Fig. 15.31 shows measured and calculated values of the base resistance and reac-
tance for a straight vertical monopole for antenna heights H less than 90°. The
monopole had a square cross-section, and h/a = 203. The characteristic impedance
of the monopole is
Z o = 60 (In ( 2 0 3 ) - 1 ) = 259 ft
The closed circles show the calculated values for the base reactance, from
Xb = —]'Z0 cot G
where G = electrical height of the monopole
= 1.05 H (assuming 5% lengthening for end effects)
VLFf LF and MF antennas 599
It is clear that the characteristic impedance calculated according to the formula
above gives results that are in good accord with experiment.

10,000

100 50 20 10 5 2
ANTENNA HEIGHT (H) DEGREES

Fig. 15.31 Base reactance —jXb and res/stance (R^ or RJ, as a function of height of the
monopole (degrees). The solid lines are measured values for a modelled antenna
over a wire-grid ground screen, the calculated reactance values (dots) show agree-
ment between theory and experiment (see text)

15.4.2 Verticalmonopoles \/4<h<5X/8


Morrison and Smith,64 Mather (private communcations) and others have measured
the impedance of straight uniform cross-section radiators (full size monopoles).
The results for these measurements are plotted in Figs. 15.32 and 15.33, showing
base resistance and base reactance versus the physical height of the radiator H in
degrees. The Morrison and Smith curve shows measurements made for a single
122 m-high antenna, of square cross-section with sides 2 m wide. The Mather curve
was derived from broadcast-station proof-of-performance measurements at some
80 stations in Canada. The closed circles, not given by Mather, have been plotted
600 VLF, LF and MF antennas

on the graph for most of the stations he used, to show the self-consistency of these
data. Since the differences between the two curves on each graph could be brought
into better agreement by shifting one of them by an amount dependent on H, the
differences are probably due to stray effects, such as base insulator capacitance,
capacitance of guy insulators, lighting transformers etc.

MORRISON
//* I
H
AND SMITH V
1
f N\
• 1

00-

' 1
\
Y
j *

r
10-

/
— i- _.
5-
40 60 80 100 120 140 160 180 200 220 240 260
PHYSICAL HEIGHT (H) DEGREES

Fig. 15.32 Measured base resistance R^ as a function of antenna height H (degrees) for
standard broadcast guyed towers of uniform cross-section

In Figs. 15.34 and 15.35 the measurements of Morrison and Smith64 are com-
pared with values calculated by two different formulations of the transmission-
line method. The curves calculated by the Schelkunoff81 method employ his
formula for the characteristic impedance of the tower, which differs from the one
adopted here. The curves labelled Belrose have been calculated by the author,
using the method given in Section 15.2.5. The relation between H and G was not
exactly the same as the one shown in Fig. 15.5, since the curves in this figure were
changed slightly after the calculations shown in the present figures were completed,
but the difference is small.
To illustrate the use of the equations given previously, let us calculate the input
impedance of a 122 m tower, having a square cross-section 2 m wide for 1300 kHz
h/X= 0.528 (H = 190°)
0 = 0.59(2) = 1.18 m
122
= 60 In -1 = 21812
1.18
VLF, LF and MF antennas 601
I000-

500-

6 200-

10-

40 60 80 I00 I20 I40 I60 I80 200 220 240 260


PHYSICAL HEIGHT (H) DEGREES

Fig. 1 5 3 3 Measured base reactance ± jX^ as a function of antenna height H (degrees) for
standard broadcast guyed towers of uniform cross-section

IUUU-

EX PERU
co 5 0 0 - BEL R O S E ^ - f ^

O SCHELKUNOFF^ J / * ^

£
/5 '?\
V
en
us 1 0 0 -
/
\
y
CO J

A
a •*> A
LJ
CO

80 100 120 140 160 180 200 250


PHYSICAL HEIGHT (H) DEGREES

Fig. 15.34 Measured and calculated base resistance as a function of antenna height H
(degrees) for a standard broadcast tower of uniform cross-section
602 VLF, LF and MF antennas

D = 3.9°
G~ 1.2(190) = 228° (see Fig. 15.5)

/*r(loop) = 15 ( — — sin 2(228) + I In — + 1.7221 cos 2(228)


\ 2 I 180
228 \
+ 4.83+ 2 I n — = 52.5 ft

From equation 15.16, the radiation resistance referred to the base of the mast is

sinh
218 ' " ~ \\ 218
?ix /
2 u2 /52.5\ 90 ft
cosh — I - (228)

Measured values for Rb (see Fig. 15.32) range from 95 to 140 ohms.

500 —

200-

80 100 120 140 160 ' 180 200 220


PHYSICAL HEIGHT (H) DEGREES

Fig. 15.35 Measured and calculated base reactance as a function of antenna height H
(degrees) for a standard broadcast tower of uniform cross-section

Equation 15.17 gives the reactance at the base


.218 sin 2(228)
VLF, LF and MF antennas 603
= —fill ohms
Measured values of Xb were about —/ 210 ohms.

15.4.3 Vertical monopoles h > 5X/8


Antenna heights greater than 5A/8 may be employed to obtain increased field
strength by dividing the tower with insulators and feeding at the place(s) where the
tower has been sectionalised. The currents in each section are maintained in phase.
Alternatively, the same effect can be achieved for a continuous grounded tower, by
employing a distribution of cages along the mast. The arrangement most commonly
used is to divide the tower into two sections of approximately one-half wavelength.
Two end-on half-wave antennas (Franklin antennas) were included in the summary
of antenna types and gains in Fig. 15.2.

(a) h = \ / 4 (d) h=X

ELEVATED FEED

Jf
"ELEVATED
FEED

BASE
SOURCE

Fig. 1536 Calculated current distribution on monopole antennas of various heights h/k for
base feed and elevated feed

An alternative arrangement for monopole heights up to h = A, is to sectionalise


the antenna at a height of 1/3 of its total height measured from the base of the
antenna (see Fig. 15.36). This Figure also shows the assumed current distribution
on the antenna for this elevated feed compared with base feed (Hatch et al.,39).
Fig. 15.37 shows the computed elevation patterns for these antennas, calculated
from the assumed current distributions on them. For h < A/2, there is little dif-
ference between elevated feed and base feed, but there is a marked difference
between the patterns for h > A/2. When h = 3A/4 the pattern for the base-fed
monopole starts to break down into a low-angle and a high-angle lobe, whereas
there is no deterioration in the pattern for the elevated-feed antenna. When h =
A, the base fed monopole has no low-angle lobe at all, only the high-angle lobe
remains; whereas for elevated feed, the pattern is squashed, exhibiting gain in the
direction of the horizon with small secondary lobes at high elevation angles.
604 VLF, LF and MF antennas

15.4.4 Folded monopole


An alternative method of feeding a grounded tower radiator is to convert it into a
folded monopole (or folded unipole). By this method, the total antenna current
is divided between two conductors which are paralleled at their current nodes, at
the top, and power is fed into one leg only; the other is grounded. The folded
unipole can be arranged as a 'coaxial' type, where the feed is between ground
and the outer sheath, a cage of wires about the grounded tower, see Fig. 15.15;

90° 70"
30*

(o)
h = X/2

0.5

(c) BASE

48.6°
30° -10°

ZZ~~~" EED ^
ELEVWED FEED
0 Q5 10

Fig. 15.37 Calculated elevation patterns for monopole antennas of various height h/k for
base feed (dashed line) and elevated feed (continuous line)

or a parallel conductor type which is shown in Fig. 15.38. This latter type has been
analysed and described in the literature [Uda and Mushiake (see Tai,92), and King
and Harrison.49]
The input impedance of a folded monopole can be calculated from expressions
given by these authors. If the two conductors are the same diameter, then
VLFf LF and MFantennas 605

where Zu = input impedance of a base driven monopole of effective diameter d


d = 2(as)m
Zt = / 2 Z c tan j3(/z + s/2)
Zc = 60 In s/a
a = conductor radius

— RADIUS a

Fig. 15.38 Folded monopole antenna

The analysis of Uda and Mushiake gives some insight into the parameters of the
expression for the input impedance of the folded monopole. They consider its
excitation as due to the superposition of two modes: a symmetrical mode charac-
terised by equal driving currents in the two conductors and an asymmetrical mode
characterised by equal and opposite currents in each arm. The former, the radiating
mode, is equivalent to a monopole of effective diameter d. The latter, which does
not lead to radiation, is equivalent to a shorted transmission line of length (h + s/2)
whose impedance is Zt.
When the electrical length of the radiator is approximately a quarter wavelength,
the input impedance of the asymmetrical mode is very large compared with the
monopole impedance;i.e.Zt > AZU and

Hence the input impedance of a A/4 folded monopole is about 144 £1.
When the electrical length h < X/4, the asymmetrical mode dominates, the
radiation resistance is very small and the antenna reactance is inductive. At
606 VLF, LF and MF antennas

frequencies between these limits, as the frequency increases to the value where the
symmetric mode becomes dominant, the reactance changes from inductive to
capacitive, and the input impedance becomes very high. Thus, the folded monopole
does not behave at all like a monopole for h < X/4.
The input impedance of a folded monopole for which s/X = 0.002, d/\ = 0.0005
has been calculated by using the computer antenna modelling code NEC described
in Section 15.2.3. The computed impedance versus h/X is given in Fig. 15.39.
These results show good agreement with values calculated from the expression given
above for ZQ. Thiele et al.9* have similarly compared calculations employing a
computer moment method (Richmond73*74) with those deduced from the trans-
mission-line model.

Fig. 15.39 Calculated input resistance and reactance of a folded monopole antenna as a
function of antenna height, for the case where sfk = 0.002, dfk = 0.0005 and
eA = 0.0025 (see Fig. 15.38)

15.5 Broadcast antennas

15.5.1 Introduction
A great deal of what we know today about vertical tower radiators, about section-
alised antennas to better control the current on the radiator, about the use of two
or more towers in directional antenna arrays, and about optimum ground systems
has been developed as a result of diligent research and experiments that have been
verified by extensive application at hundreds of MF broadcast stations. In addition,
VLF, LF and MF antennas 607

there are special requirements for broadcasting in the LF band, where an adequate
bandwidth must be achieved without sacrificing antenna radiation efficiency. And,
with the increased development of the broadcast band, and the higher powers
currently employed, development of directional antenna arrays that have ever
increasing requirements on the depth of the null of the directional array, and on the
stability of the directional pattern, has necessarily led to investigations of re-
radiation from various conducting metal structures that are located nearby and
which could cause a distortion of the directional pattern. In spite of the fact that
the early development of broadcast antennas began in the mid-1920s (Ballantine,6'7)
and that our understanding of the modern broadcast tower radiator had evolved
to its present state by about 1934, research and development is still being carried
out to optimise the development of ever increasing broadcast services within the
limited frequency spectrum. Some of these aspects of antenna design will be
described in this Section.
Economics also influences antenna design. For example, the high cost of tall
towers has resulted in the requirement to diplex two broadcast stations operating
on adjacent frequencies into the same tower radiator. The high cost of land, since
land having good conductivity and flat terrain contours is usually the same land
that is ideal for farming, has resulted in the need to place more than one directional
array, operating on nearby frequencies, on the same antenna site. There are also
special requirements to change the pattern of a particular station from a day to
a night pattern to protect listeners in a nearby broadcast service area. Some of
these more technical details will be discussed in a later Section.

15.5.2 LF broadcast antennas


Transmission bandwidth and radiation efficiency are the two more important
parameters influencing antenna design for LF broadcasting application. Unfortu-
nately, for short antennas, it is difficult to achieve both together, since high effi-
ciency usually implies narrow bandwidth. This problem is especially critical at
frequencies near the lower frequency limit of the LF broadcasting band (150 kHz).
A conventional vertical radiator for 155 kHz would have to be some 450 m in
height in order to ensure sufficient bandwidth and high efficiency. A 200 m mast
having a square cross-section 2 m on a side would have an input impedance:
ZQ = 5 - / 3 0 1 ohms
Obviously some form of top-loading would have to be employed, but with con-
ventional top-loading it would not be possible to achieve the required bandwidth
and high radiation efficiency desired.
Two approaches can be taken to achieve the required bandwidth: increase the
diameter of the radiator, or multiple-tune several vertical towers or down-
leads sharing a common top-hat. Both approaches have been taken in the case of a
broadcast antenna installed near Donebach/Odenwald, Germany (CCIR 1966-69
Doc. X57). The antenna is described as a folded monopole, but since its effective
electrical diameter is about 106°, and since four towers are tuned, it is in effect a
608 VLF, LF and MF antennas

type of multiple-tuned antenna system. It comprises four 200 m guyed towers


(h/X = 0.1) all of which are insulated from the ground. They are set up in such a
way that three towers are uniformly spaced about the centre tower, at distances
of 330 m. Each of the four towers has its own star-like radial ground system, which
has a radius of 200 m. The tops of the towers are joined together by a flat wire
system 10 m wide. The inductive reactances used to terminate the outer towers
were made equal to the capacitive component of the input impedance at the
centre tower. With each tower terminated in /95 ohms, the input impedance of
of the centre tower is
ZQ = 7 3 - / 9 7 ohms
The radiation efficiency achieved exceeds 90%, and the bandwidth is 6 kHz.
It is interesting to note that the antenna resistance (largely radiation resistance)
is almost increased by a factor N2 over that for a single tower (N = number of
towers).
Recall that with multiple tuning, or for that matter, according to the theory of
folded monopoles,
Rrr = RrN2
Xbb = XbN
where Rr and Xh are the impedance of the equivalent radiator, which cannot be
simply calculated since its equivalent electrical diameter is large.

15.5.3 Anti-fading antennas


Rather elaborate collinear antenna arrays have been installed at some MF broadcast
stations to increase the horizontal gain, and reduce radiation in directions other
than the horizon, especially at elevation angles which would lead to night-time
skywave propagation to distances within the useful ground-wave service area. From
the experience in operating these antenna systems, it is evident that a high-efficiency
anti-fading antenna should be sectionalised, so that the current on it can be con-
trolled, and it must have a height of at least 2X/3 to A, to produce the ground-
wave coverage desired and the necessary rapid rise of skywave field strength for
distances beyond those of the normal ground-wave service area. If the current
distribution on the lower section can be controlled with respect to the current
on the top section, this in effect permits such a tower to possess a family of vertical-
radiation characteristics (Jeffers,44 Page and Monteath,68).
In the realisations achieved to date, either the mast was divided into two or
more mutually insulated sections (cf. Page and Monteath,68 Head,40 Knight,51)
or the mast was screened by a distribution of cages along it, and the cages become
the radiator (cf. Aizenberg et al.2 Larcharnay54'55).
Knight50 has given formula for the screening factor and characteristic impedance
of a tower surrounded by a cylindrical cage of wires. The screening factor s is
defined as the current on the outer conductor (the cage) to the total current, and
therefore s = 1 corresponds to a perfect screen. He has considered a square tower
VLF, LF and MF antennas 609

of side width 1.8 m surrounded by a cylindrical cage of radius 1.8 m. If the cage
consists of 32 wires each 0.625 cm diameter, Knight has calculated that 5 = 0.85;
hence the power radiated by the tower is 15 dB below the power radiated by the
cage. Belousov and Kouzmetsov8 used 18 wires in the cage antenna described by
them. Bern used even fewer wires in the cage, and reports a screening factor very
much higher than calculated by Knight. According to a summary of his experiments
(CCIR Doc. 1970-73, Doc. 10/62), he reported that, for a cage of nine wires,
the screening factor was 0.97. While this difference is not resolved here, it is clear
that a cage with relatively few wires will provide satisfactory performance, and
there is, therefore, no fundamental difficulty involved in driving a base-earthed
mast, or in sectionalising a mast without physically doing so. The cage of wires
must be physically bonded together at the top and bottom of the cage by a peripheral
wire. The cage type of radiator has seen greater application because sectionalising
the mast by means of an insulator involves serious mechanical/electrical difficulties
and expense.
A A/4 stub can also be used to sectionalise a mast, because it effectively reflects
a very high impedance into the mast at the place where the stub is open. A quarter-
wave stub can be used to match a A/2 radiator, and effectively isolate its driven end
from the ground beneath it and the transmission line feeding it (the J-antenna);
or it can be used to reduce the re-radiation from unused towers affecting the
directional pattern of broadcast antennas. Both these applications are further
discussed below.

fin Iff
1

Fig. 15.40 Various methods of feeding grounded tower antennas

Fig. 15.40 illustrates various arrangements for feeding and sectionalising grounded
towers. In (a) the cage is used to feed a A/2 grounded tower (Knight,50). In this cir-
cumstance, the cage (labelled 1) is bonded to the mast (labelled 2) at its top, and the
feed is between the bottom of the cage (3) and ground. The sketch also shows the
equivalent circuit. In the arrangement at (c), an integrated A/4 matching stub is
610 VLF, LF and MF an tennas

employed, to end feed a X/2 radiator. This arrangement is popularly called a J-


antenna. The balanced feeder is tapped onto the stub mast and the tower at a
place where the impedance of the antenna equals the impedance of the feeder.
This arrangement effectively end-feeds the X/2 radiator, which, insofar as gain
and pattern are concerned, is insulated from ground and elevated to a height of
X/4 between its bottom end and ground. The gain and antenna pattern are illustrated
in Fig. 15.2.
The arrangements in Fig. 15.405 and d are cage-type radiators used to section-
alise a tower. If the bottom cage in (b) is grounded to the tower at its upper end,
and insulated from the tower at its lower end, the radiator can be conveniently
fed by a coaxial cable, which is routed up inside the cage, its sheath is connected
to the upper end of the bottom cage and its inner connector to the lower end of the
top cage, which is insulated from the tower (Larchamay54). The upper end of the top
cage can be insulated from or connected to the mast, since the current is zero there.
The arrangement in (d) is similar to (b) in regard to physical connections to the
tower, and methods of feed, except that the effective height of the radiator is now
X instead of X/2. Since there are two feedpoints, the design engineer has a wide
range of adjustment. If the upper and lower radiators are fed with equal powers
and in-phase currents, the gain and pattern obtained are that of the Franklin
antenna, two end-on half-wave antennas in phase (Fig. 15.2).

Fig. 15.41 Method of feeding and equivalent circuit for a grounded tower cage antenna,
sectionaUsed into three equal lengths

One of the most complex antennas so far utilised is that described by Belousov
and Kouznetsov,8 shown in Fig. 15.41. It is a cage type; the antenna is effectively
sectionalised into three sections, two of which are fed; and a variable reactance
VLF, LF and MF antennas 611

(physically lengths of coaxial cable) permits control of the current on the lower
cage. The upper and central antenna sections may be considered as a symmetrical
dipole with arms H/3 in length. The lower section of height H/3 is an asymmetrical
dipole. The mast height was 321 m. The authors describe the antenna and show
calculated and measured vertical patterns at various frequencies across the MF
broadcast band.

15.5.4 Directional arrays


The theory and design of MF directional antenna arrays is a specialised field that
has required considerable mathematical computation and, above all, practical
experience to become proficient. A detailed description of design procedures
is beyond the scope of this book. The basic concepts are, however, simple and
straightforward, since the total field at any point is calculated from the vector
sum of the fields at that point due to each tower in the antenna array, which is
given by

E = I EKfK (6) (fa


K— 1

where EK is the magnitude of the field intensity at unit distance in the horizontal
plane produced by the kth tower, fK(6) is the vertical pattern of the kth tower,
and n is the number of towers. The phase factor $K is usually expressed as the sum
of a space phasing, which results because of the location of the kth tower with
respect to a vector reference axis; and a time phasing, which is the electrical phase
angle measured with respect to this same axis, due to the particular phase of the
current on the kth. tower.
The early antenna patterns for 2-tower arrays were hand calculated. Later, in
the 1940s a systemised approach was taken, and hundreds of antenna patterns for
2- and 3-tower arrays were tabulated (Smith,85). These calculations were carried
out on a simple electromechanical calculator.
Nowadays, with the advent of minicomputers, most broadcast engineers have
computer programs which will calculate the resulting field due to a general con-
figuration of any number of towers, each with its own time-phase vector relation,
since the vector sum given by the above relation is basically a very simple cal-
culation. However, in the design of an antenna array to meet certain specific
requirements in regard to its directional pattern, a synthesis approach is needed;
that is, the pattern requirement is known, and what the designer engineer must do
is design an antenna array that will produce the required pattern. Since the numeri-
cal methods for pattern synthesis have not so far been developed, and since it is
not possible by examination of the general vector equation given above to decide
how to proceed, it is at this point that design experience is required. Many direc-
tional antenna arrays have four towers at the corners of a parallelogram. The
chief reason for the popularity of this array is the relative ease with which the
pattern shape can be designed to meet complicated requirements. A 4-tower
parallelogram can be considered as two sets of 2-towers, the resulting pattern is
612 VLF, LF and MF antennas

then calculated by pattern multiplication. Similarly, six towers can be treated as


three sets of 2-tower arrays. Since patterns for 2-tower arrays have been tabulated,
the design engineer can start out his design by ensuring that the pattern nulls will
appear in the desired directions. Having then arrived at a best-guess antenna array
configuration, the designer can then utilize a computer program to solve the gen-
eralised equation above, and, by iteration, that is randomly moving the towers
by small amounts, or the phases of the currents on them, he can achieve a best fit
to the desired pattern.
The calculation is at best only an approximation because the current distribution
is not exactly sinusoidal or in phase along the tower, and no account is taken of
base insulator capacitance, effects of guys etc. The final pattern is therefore estab-
lished by field measurement, which in itself requires considerable experience
(Smith,86). In fact, 'taming' the MF directional array is the title of one of the articles
in the above-referenced publication. Sometimes problems arise such as unwanted
pattern nulls etc. In some directional antenna systems, a tower will have a negative
resistance at the driving point. This means that power must be removed from the
tower terminals. While the initial adjustments are made, this power can be dissipated,
but, in the final operation, it is desirable to feed this power back into the system to
maintain high efficiency. Also, the so-called optimum design takes no account of
tower voltage, and some towers in a complicated antenna array may have rather
high voltages on them, especially if they are not radiating much power (Rb ~ 0)
and the current and reactance are high. This has sometimes resulted in additional
complications.
Obviously, with the present availability of microprocessors and minicomputers,
more attention could be given to proper pattern synthesis, and, by employing
large computers and sophisticated antenna modelling codes (moment methods),
the pattern could be more rigorously calculated, and the calculations would provide
also the effective base impedances of all towers in the multi-tower array, and the
actual current distribution on the towers. Thus, in an actual antenna array, the
calculated currents in amplitude and phase should compare more exactly to those
measured, thus facilitating initial set-up and reducing the time required to achieve
the desired pattern by field measurement.
To illustrate some of the pertinent factors in antenna design, and parameters
relevant to these designs, two antenna arrangements will be considered. In one,
a 3-tower array is used to provide a wide deep null in a specific direction; in another,
a critical array is described that provides protection in several directions, and
extreme protection in one direction.
The CBC station CHFA, operating on 680 kHz, located near Edmonton,
employs three towers, equally spaced on an aximuth 10.5° east of north. The
spacing between the towers is 1103 m (90°). The towers are all 66.8 m high (70.9°),
and the south, centre and north towers are fed with relative currents of 0.538/97.5°,
1/0° and 0.484/—97.5°. This arrangement provides a pattern with a null over
a wide range of angles (80°). The calculated and measured patterns are shown
in Fig. 15.42; the radiation pattern in (b) gives detail of the null on an expanded
VLF, LF and MF antennas 613
330° VT 20° 30°

\
40°

130 140 150 160 170 180 190 200 210 220 230 240 250
AZIMUTH • DEGREES (b)

Fig. 15.42 Measured (dashed line) and theoretical (continuous line) horizontal radiation
pattern, and part of the pattern on an expanded plot, for MF broadcasting station
CHFA (680 kHz), Edmonton, Alberta. The antenna system employs three towers,
equally spaced on a line bearing 10.5° (after CBC, 1976)
614 VLF, LF and MF antennas

scale. This is an example of a simple design showing good agreement between


theory and measurement. The omnidirectional pattern for lOkW transmitter
power is predicted to be 593 mV/m at 1 mile (954 mV/m at 1 km), and therefore
the field strengths in the main beam and in the null directions are + 4 dB and
— 35 dB respectively.
The design for the proposed new antenna array for MF-AM broadcasting station
CFTR, operating on 680 kHz at a proposed power level of 50 kW, is an interesting
example of a critical array (Rogers Radio Broadcasting Ltd., Toronto). The trans-
mitting antenna array will be located near Grimsby, Ontario, on the south shore of
Lake Ontario, with its main beam directed north towards Toronto, and the pattern
null directed towards Rochester, NY, about 160 km east of Grimsby. Station
WNYR, Rochester, operates on the same frequency with an omnidirectional antenna
and a transmitter power of 0.25 kW. The field strength of CFTR over an arc of
about 20° in the direction of Rochester must not exceed 4 mV/m at 1 mile (6.4
mV/m at 1 km). The design is an 8-tower parallelogram array, employing towers
122 m (99.5°) high. Each tower will have its own radial ground system, of mini-
mum length 0.25X. The outer radials will run to the edges of the antenna site,
which is a wedge-shaped piece of land (Fig. 15.43a) having a total antenna area
equivalent to about 0.4A radius. The radials will terminate in Lake Ontario along
one long side of the parallelogram array. A 30 m diameter elevated counterpoise
(the open circles on the site sketch) beneath each tower is proposed for purposes
of increasing the stability of the pattern. The predicted omnidirectional RMS
field strength for a transmitter power of 50 kW is 1.2 V/m at 1 mile (2.03 V/m
at 1 km). The theoretical radiation pattern of this proposed array is shown in
Fig. 15.435. Note that the predicted field strength in the main beam is + %.% dB,
and in the null is — 58 dB with respect to this RMS field. The design has also
considered pattern stability, and the median field in the null is estimated to be
3.5 mV/m at 1 mile (5.6 mV/m at 1 km), which is - 51 dB below the RMS field.
The predicted array operating parameters are given in the Table below.

Tower Base operating Base Power


no. impedance (ohms) current (amperes) (watts)
1 1.4+/18.1 3.8 /168.7° 19.8
2 24.4-/1.2 11.1 /129.4° 3019.7
3 23.6-/33.4 11.1 /90.1° 2920.7
4 -55.3-/87.8 3.8 /50.8° -781.9
5 226.6-/53.6 3.8 /320.7° 3203.8
6 179.4+/50.5 11.1 /O0l 22202.4
7 143.8 +/51.1 11.1 /39.3° 17796.6
8 114.5 +/3.2 3.8/78.6° 1618.9
50000.0
VLF, LF and MF antennas 615

Lakeshore

Notes:
Total effective area equivalent to 0.4.A radius
Minimum radial length to be O.25X
Dimensions are approximate
— Limit of ground system
Q 100' diameter elevated counterpoise

Fig. 15.43a Tower configuration and ground system layout of MF broadcasting station
CFTR (680 kHz) at its proposed new site near Grimsby, Ontario. The open
circles are the limits of the 30 m diameter elevated couterpoises

It is interesting to note that 80% of the transmitter power is radiated by towers


6 and 7, and that the resistive component of the base operating impedances range
from + 227 £2, through near zero (+ 1.4 £1) to — 55 £1.

15.5.5 Distortion ofMF broadcasting antenna patterns due to re-radiation


Many MF broadcasting stations employ directional arrays that have been designed
to a close tolerance with respect to permissible levels of radiation in particular
directions, which could cause interference to listeners in an adjacent broadcast
area. With population growth and civilisation encroachment on once isolated
transmitting sites, re-radiation from towers, tall buildings, power lines and other
structures can result in pattern distortion, and in difficulty with maintaining
an adequate null in the directional pattern. If the re-radiator is extended and
616 VLF, LF and MF antennas

320°
330° 340° 350° (
) 10° 20° 30°

N
40°

<

/
X \
310° 50°
\
/ \
300°
> <

290° 70°

280°

t Ml
270° 90°

260° .——• •—.100°

250° 240° 230° 220° 200° 180° 160° 140° 130° 120° 110°

35 -|

)H 15 -
u

S' MEDIAN
/ .RESULTANT
FIELD

80 90 100 110 130


AZIMUTH (DEGREES)

Fig. 1543b Theoretical horizontal radiation pattern, and part of the pattern on an expanded
plot, for MF broadcasting station CFTR (680kHz), at its proposed new site,
near Frimsby, Ontario. (After Rogers Radio Broadcasting, 1980)
VLF, LF and MF antennas 617
physically large, such as an EHV (500 kV) power line, and is close to the trans-
mitting antenna, even the main beam can be adversely affected.
High-rise buildings, having cross-sectional dimensions of 30 x 30 m, and heights
of 60 m are resonant in the middle of the MF broadcast band. EHV power lines
are particularly troublesome. It just so happens that 500 kV power lines which
employ 54 m towers spaced 274 m, with a 'skywire' for purposes of 'grounding'
the power line for lightning protection connecting the tops of the towers, are
2-wavelength loop-resonant at 860 kHz. That is, the loop consisting of a tower,
the skywire to the adjacent tower, that tower and the image of this half-loop in
the ground has a perimeter of two wavelengths and is resonant. A power line
comprising a long row of evenly spaced towers acts like a series of over-coupled
resonant circuits, absorbing power from the broadcast antenna array and re-radiating
it.
A number of researchers have discussed re-radiation effects from tall TV towers
and other structures, e.g. Presthold,69 Alford and French,4 Flood and Field.32
More detailed attention has been given to measuring the magnitude of these effects
and to developing predictive techniques by Belrose et al.15 and a great deal has
been learned. The understanding of the re-radiation mechanism was first sought by
computation and model measurement of re-radiation by simple structures such as
thin towers and straight transmission lines with uniformly spaced towers. The
numerical analysis employed thin wire models for analysis by prepackaged computer
codes such as NEC (Burke24). The values of key parameters of the model were
formed by a combination of variation studies and theoretical guidelines, such that
agreement was obtained with the scale-model measurements on an outdoor ground-
level antenna pattern range. Both the azimuthal pattern and the frequency depen-
dence were studied (Trueman et al.9*). Also, the effect of detuning stubs on the
various structures to minimise the re-radiation has been studied (Belrose et al.16
Trueman et al.99 Royer76).
Figs. 15.44 and 15.45 show calculated values for the normalised scattering
cross-section for grounded and ungrounded towers of various heights and thickness.
The scattering cross-section (o) is a quantitative measure of the power density at
a given range from the re-radiator, in terms of the power density of the signal
incident upon it. The computer program NEC provides phase as well as amplitude
for the scattered field, but the amplitude gives limits on the magnitude of the
effect, which is sufficient for the purpose of deciding whether a re-radiator will
have a significant effect on the pattern of the broadcast array. Royer76 has shown
that these limits are:
,^V2 /G a (18O>(0)V /2 l 2
G(0)
1/2] 2

where
G a (0) = the gain of the array in the direction 0 when the tower
re-radiator is not present
618 VLF, LF and MF antennas

(j> = direction shown in Fig. 15.46.


rt = distance between the array and the tower
Ga(180°) = gain of the array in the direction of the tower
G(0) = total gain in the direction 0, of the array and the tower
combination
h = height of tower
a = radius of tower
X = wavelength of transmitted signal
= the scattering cross-section of the tower

Fig. 15.44 Normalised scatter cross-section for a grounded tower as a function of tower
height and thickness

And

For the cardioid pattern shown in Fig. 15.46, which is produced by two X/4
monopoles spaced X/4 and with the current in monopole 1 leading that of mono-
pole 2 by 90°, Ga (0) is given by
VLF, LF and MF antennas 619
2
Ga(<p) = 6.56cos {45° [1 - c o s (0 - 150°)]}
Then if X = 300 m (i.e./= 1000 kHz)
h = 70.5 m; h/\ = 0.235
a = 2.4 m; a/X = 0.008
rt = 1500 m
it follows from Fig. 15.43 that
2
a /X = 0.85

Fig. 15.45 Normalised scatter cross-section for an insulated tower as a function of tower
height and thickness

Therefore,
o = 0.85(300)2 = 7.65 x 104 m2

and = ! 75S x
2
2
4TT(1500)
The total gain of the array, in the direction where previously there was a null,
is therefore
620 VLF, LF and MF antennas

G(-30°) = 1.755 x 10~2


This puts G ( - 30°) 25.7 dB below the peak value Ga(150°)
When the structure is ungrounded, Fig. 15.44 shows that
a/X2 = 0.033
which is 14.1 dB below that for the grounded tower; hence
G(-30°) = - 2 5 . 7 + (-14.1) = - 3 9 . 8 d B

TO FAR FIELD POINT

t -TOWER

GAIN PATTERN FOR


-f>--30°

ARRAY WHEN NO
TOWER IS PRESENT

Fig. 15.46 A tower affecting the radiation pattern of an MF broadcasting antenna

From this calculation, we can estimate that installation of a A/4 stub on a


grounded tower, with the shorted end towards the top of the tower and the open-
circuited end facing down, near the base of the tower, would be approximately
equivalent to 'lifting' a grounded tower off the ground plane. This has in fact been
confirmed by calculation and experiment.
When the height of the tower h is greater or less than X/4, it is still possible to
minimise the re-radiation by installing a X/4 stub or a full-length stub tuned by a
suitable reactance at its open end. The curve in Fig. 15.47 shows the variation of
the effective normalised scattering cross-section as the reactance across the open
end of the stub is 'tuned' for a tower 0.4X high.
Isolated high-rise buildings behave like thick towers. Compared with the scatter-
ing by thin towers, the height for X/4 resonance is shorter and the decrease in
scatter cross-section between X/4 and X/2 is reduced (see Fig. 15.48).
Power lines are much more complicated, and since the scatter from the power-
line is not omnidirectional, phase as well as amplitude must be calculated to predict
the re-radiation effect. To illustrate the magnitude of the effects, Fig. 15.49 illus-
trates a real situation (simplified model), where a 500 kV powerline (dimensions
given on the Figure) runs near the broadcasting stations CBL/CJBC at Hornby,
Ontario, operating on 740 and 860 kHz. The transmitting antenna is a 196 m high
VLF, LF and MF antennas 621

tower (approximately a 5X/8 radiator). The power line runs past the radiator, and
near to it, 1470 ft (448 m) about 1.2A distant. The results in Fig. 15.50 show com-
puted and measured azimuthal patterns (at a scale model frequency) for 515 and
860 kHz. The powerline is approximately resonant at these frequencies. In the
absence of the powerline, the azimuthal pattern would have been a perfect circle.

XY PLANE PERFECTLY
CONDUCTING

TO FAR-FIELD POINT
WHERE CTfl(0,(?)/A2
IS COMPUTED
INCIDENT PLANE
"WAVE
-•X

XL(OHMS)

Fig, 15.47 Scattering cross-section of a grounded tower with a tuned stub

A powerline can be detuned by installing detuning stubs on the skywire, as


shown in Fig. 15.51. A modelled powerline consisting of nine towers, like Fig.
15.49 but extended, has been successfully detuned, see Fig. 15.52. The max./min.
ratio, a measure of the scalloping of the omnidirectional pattern, as a function of
frequency for this 'powerline' of nine towers, is shown for no detuning stubs and
with one stub on each skywire tuned to 860 kHz (Belrose et al.16).
Large structures such as radio masts, buildings and power transmission lines
can be resonant at MF frequencies; when they are, such structures re-radiate
622 VLF, LF and MF antennas

-H !
— a o » WIRE RADIUS

INC. PLANE WAVE

(a) (b)

a o /A «0.001

*-«--f « t -*

WIRE GRID AS SHOWN IN (a) ABOVE


X 9 =0°
••• 9 =180°

WIRE GRID AS SHOWN IN (b) ABOVE


0.05-
9 9= 0°
e <P= 180°

T
0.1
r i
0.3
I
0.5
0.2 h/A

Fig. 15.48 Scatter cross-section for a wire grid building 0.1\ on a side as a function of the
height of the building

Fig. 15.49 A 500 kV power line adjacent to an omnidirectional monopole radiator


VLF, LF and MF antennas 623
1 0 DFTTBKT, SCALE

(a)

515 kHz

Comparison of
oonputations and
measurements.

(b)
860 kHz

Fig. 15.50 Theoretical and measured (continuous line) azimuthal patterns for an omnidi-
rectional antenna affected by a power line (comprising 9 towers): (a) 515 kHz and
(b) 860 kHz. The physical arrangement is illustrated in Fig. 15.49, and the fre-
quencies correspond to the 1\ and 2\ loop resonant frequencies of the power line.

^—tower

tower

Fig. 15.51 A method of detuning a power line by installation of a stub on the sky wire
624 VLF, LF and MF antennas

strongly, and this re-radiation can distort significantly the pattern of a nearby
broadcasting antenna array. The work to date (numerical and experimental modell-
ing) has considered the case of a perfectly conducting ground plane. The analysis
must be extended to include losses due a finitely conducting real earth. Also, while
initial studies using detuning stubs have shown that the re-radiation can be signif-
icantly reduced, the installation of such stubs on powerline skywires has not yet
been accepted as structurally satisfactory from the power utilities' point of view,
and the installation of detuning stubs on highrise buildings presents even greater
difficulty.

1 1 1 1

10 -
no detuning stub v \
8 IP* ~°~ $ -
measured — - S " " ^ >^
calculated-> / V
6 V -
/
J
/
V
4 / \
/ ybent detuning stub
2 _ 1/1/^-straight detuning stub _

i
750 800 850 900
FREQUENCY in kHz

Fig. 15.52 Calculated and measured scalloping of an omnidirectional pattern caused by a


9-tower power line with and without stub tuning on each sky wire (the stubs
were tuned to 860 kHz)

Sather77 has discussed the use of tuning stubs on the broadcast radiators them-
selves to reduce currents on them due to nearby broadcast stations.

15.6 Portable and mobile antennas

While it is technically possible by employing a team of men and suitable equipment


to erect breakdown-tower antennas, or to erect towers that support wire antennas,
for example for military communications, these are transportable not portable
antennas. For portable application employing simple wire antennas, some means of
support other than a transportable tower must be employed. This could be a high
cliff, a kite or a kitoon (a combination between a balloon and a kite). If feasible,
the antenna should be operated at half-wave resonance to avoid the need for a
VLFf LF and MF antennas 625
13
ground system, unless an ocean can be used for the ground plane (Belrose ). In
areas where the ground is a poor conductor, such as arctic terrain, half-wave dipole
antennas which are strung out over the surface of the ground can be employed
(Evans30). Wire antennas can also be deployed from an aircraft or held aloft by a
helicopter. In the former application, a counterpoise wire is usually employed, so
the 'antenna' can be fed against a conductor that is physically much longer than the
body of the aircraft. This is particularly necessary for operation at the lowest
frequencies (Franchi34).
For marine and land-mobile applications at MF and lower frequencies, practical
antennas are electrically short and the antenna must be 'loaded' in some way
(usually inductively) to increase the current area on it and decrease its capacitive
reactance. In such mobile application, the antenna is operated under conditions
where its electrical properties (current distribution, impedance characteristics,
radiation pattern and efficiency) are strongly affected by the platform environ-
ments; i.e. the structures on which the antennas are mounted, such as automobiles,
tanks, wooden or fibreglass ships, etc. While ungrounded antennas with elevated
feed (Niver and Hizal65) can minimise the effect of the platform, electrically short
antennas must be fed against the platform to achieve reasonable radiation. In some
cases, a stake is driven into the ground to provide an earth terminal or a few radials
are laid on the surface of the ground; this not only results in a rather imperfect
ground system, but can introduce directional effects.

15.6.1 VLF/LF aircraft wire antennas


Fig. 15.53 shows a VLF/LF trailing wire antenna. A shorter 'counterpoise' wire is
employed so that the wire 'antenna' can be fed against a conductor that is elec-
trically longer than the body of the aircraft. In this 2-wire antenna system, the
shorter wire trails a stabilising drogue and so it is essentially horizontal, whereas the
longer wire has an aerodynamic weight. The sketch shows the aircraft flying on a
straight course, but if it were to fly in a specified circular pattern this would allow
the longer weighted wire to drop vertically, thus greatly increasing the vertically
polarised component of the radiated power. In practical application, the role
usually played by the antenna and its 'counterpoise' are reversed. That is, the
effective feeding point is at the aft of the aircraft where the shorter wire exits
from the aircraft, the longer wire being grounded to the aircraft skin (Martin et al.62).
Lengths of the counterpoise wire of 0.08 to 1.4 km and lengths of the long wire
of 0.5 to 8.5 km have been employed for operation in the frequency band 17-300
kHz. Obviously, there are severe technical difficulties in deploying such long wire
antennas and the speed and direction of the aircraft affect its radiation charac-
teristics (Kossey52). The evolution of the antenna and of others tried is briefly
described by Franchi.34

15.6.2 Dipole antennas on the surface of the ground


A horizontal dipole lying on the surface of the ground, or suspended above it by a
few metres, makes a rather inefficient, but effective antenna for portable application;
626 VLF, LF and MF antennas

the vertically-polarised field component is a maximum along the axis of the antenna.
Since the efficiency of the radiator is inversely proportional to the ground con-
ductivity, the antenna is only suitable for use in locations of very poor ground con-
ductivity such as frozen arctic tundra, old lava beds and arctic or antarctic ice caps.
Long wire horizontal dipoles have been used in Hawaii (Seeley and Wiborg79) for
VLF frequencies 8-24 kHz, in the antarctic for frequencies 5-14 kHz (Raghuram
et al10) and for a portable LF communication system for use in the Canadian
arctic (Evans30). It is interesting to note that, although the antarctic antenna was
above an ice cap of very low conductivity, nevertheless it was necessary to mount it
on wooden poles to prevent it melting into the ice. For half-wave resonance, the
efficiency of the antenna at Siple, Antarctica, was estimated to be about 2% over
the frequency range 5-14 kHz.

Fig. 1553 VLF/LF trailing wire antenna and counterpoise wire for aircraft application

Evans30 has applied a rather simple expression for the efficiency of dipole on
the ground, an expression given by Weeks and Fenwick,103 which, while not exact,
seems to give agreement with experiment:

144
V =
R. 1-F2 \ cos —
2
where T? = efficiency defined as the power density at the receiving location
when the dipole antenna is transmitting divided by the power den-
sity at the receiver location when a quarter-wave vertical is trans-
mitting
Rs = input series resistance

n = the complex index of refraction of the earth = jer —/-

_ measured resonant frequency


F —
resonant frequency in air
o = ground conductivity
er = relative dielectric constant of the earth
VLF, LF and MF antennas 627

e0 = dielectric constant of free space


Cx> = angular frequency of transmission
For practical ground conditions

coe0

Accurate field strength versus distance measurements for a 610 m dipole showed
that o = 0.5 mS/m, the antenna was resonant at 200 kHz, and the measured effi-
ciency was —20.4 dB. The measured parameters to apply the above formula are:
Rs = 154 £1 and F= 0.61. Thus, the calculated theoretical efficiency is - 2 1 . 9 dB
(about 1%). These measurements were made for a dry sandy soil. Arctic tundra
has a similar conductivity (0.3 mS/m), while glacial ice has a conductivity an order
of magnitude smaller. Since antenna efficiency is inversely proportional to con-
ductivity, radiation effiencies of 10% would be expected for this situation.

15.6.3 Electrically short mobile whip antennas


Except for the larger ships, where a T- or L-type antenna can be employed, sus-
pended between the ship's masts, MF marine and land-mobile antennas are typically
short vertical monopoles or whip antennas. Since these antennas are electrically
short, they are usually inductively loaded. Fig. 15.54 shows two types of inductively
loaded whip antennas. In (b) the loading is by a lumped inductance; in (c) the
antenna is loaded by a distributed inductance and, in effect, the antenna is a helix
whose diameter is electrically small. These antenna types should be compared
with the conventional base-loaded antenna shown in Fig. 15.5%.

llo

wwwwwwww wwwwwwwwww wwwwwwwww


a) c)

Fig. 15.54 Electrically short monopoles; (a) Base-loaded; (b) Inductively-loaded; (c) Con-
tinuously loaded

Hansen37 and Fournier and Pomerleau33 have analysed inductively loaded


antennas of the type shown in Fig. 15.542? by numerical modelling moment method
techniques. The latter authors also experimentally measured the current distri-
bution on modelled antennas. Much earlier Belrose9 analysed such antennas by the
628 VLF, LF and MF antennas

simpler equivalent transmission-line method. This early analysis assumed linear


current on the antenna. The following is a modification of that analysis for an
assumed sinusoidal current distribution.
The inductance Lo is adjusted so that the base input impedance of the centre-
loaded whip is purely resistive, that is, the antenna is resonant. The input impedance
of the antenna can therefore be calculated in the manner previously described. The
lower section of the antenna is considered to be a transmission line loaded by the
capacitive reactance of the top section of the antenna with the inductive reactance
of the inductor in series. For resonance,
XL = jZ0 (cot Gx ~ tan G2) ohms
where Gt and G2 are the electrical lengths of the top and bottom sections of the
antenna.
The radiation resistance of the antenna can be calculated by adding the current
areas on the two parts of the antenna since the currents are in phase.
180/l-cosGA
y^i — — cos u1! degree-amperes
7T y sin Qj\ I
180
A2 — sin G2 degree-amperes
n
If h/X = 0.05
h/a = 500

Z o = 60 (in 11 = 313 ohms

Hence if
Hx = H2 = 0.05(360)/2 = 9°
and d = G2 = 1.05(9) = 9.45°
the inductance Lo for base resonance is
XLQ = /1828 ohms
37
Hansen calculated XLQ — 1823 £1 (as read off graphs given in his paper). Thus
A 2 = 9.41 degree-amperes
Ax = 4.67 degree-amperes
and Rr = 0.01215(4! + A2f = 2.41 £1
If the coil Q-factor were 300, the radiation efficiency for Rg = 0 is calculated to
be 28% (Hansen calculated - 25%).
These operational parameters can be compared with conventional base loading
(Fig. 15.54a). Here
VLF, LF and MF antennas 629
Xa = -J'Z0 cot Ga = - / 9 1 4 ohms
180 / l - c o s G a \
A = 1 = 9.54 degree-amperes
7T \ s inG a /
Rr = 0.01215 A2 = 1.1 ohms
and the radiation efficiency for a coil Q-factor equal to 300 for Rg = 0 is calculated
to be 26%.
Let us now consider the helically loaded antenna, illustrated in Fig. 15.54c.
Kandoian and Sichak47 have analysed in some detail tuned helical antennas. If a
length of wire approximately a quarter-wavelength long is wound onto an insulated
flbreglass whip, the number of turns and the length of the wire can be adjusted
so that the input impedance of the whip is purely resistive. The antenna is then
'quarter-wave' resonant. Clearly, for a helical antenna the velocity of propagation
down it, v, is much less than the velocity of light, c. The radiation resistance can be
calculated approximately from the expression

Rr = l^\ Rr(\/4)

where Rr(K/4) is the radiation resistance of a quarter-wave whip, equal to 36 £2.


For quarter-wave resonance, this velocity with respect to the length of the antenna,
h, must be related by

h/\ =

and solving for v/c and substituting in the equation above,

For h/X = 0.05, Rr= 1.44 ohms.


This value lies between the values calculated above for base- and centre-loading.
Since the Q-factor for a helix can be the same or better than the (Q-factor for the
centre-loaded whip, its radiation efficiency can be at least as good. The operating
frequency of the helical antenna is not, however, easily changed, but it can be
readily matched by grounding it and tapping the feed point up the appropriate
distance to match the coaxial feeder cable. The base-tuned whip can, however,
be similarly matched and its operating frequency is more readily changed.

15.6.4 Interaction of antennas with their platforms


Mobile antennas are usually operated under conditions where their electrical
properties (current distribution, impedance characteristics, radiation pattern
and efficiency) are strongly affected by their platforms; i.e. the structures on
which they are mounted such as automobiles, tanks, wooden or flbreglass ships,
etc. The frame of an automobile at medium frequencies is an inadequate ground
630 VLF, LF and MF antennas

plane, because its physical size is a small fraction of a wavelength, and it is insulated
from the ground beneath by the rubber tyres of the vehicle. Track vehicles are
'grounded' but imperfectly through a sliding contact with the track, and the track
in some cases may be rubber. Metal ships of large physical size provide an adequate
ground plane to feed the antenna against, but wooden or fibreglass ships pose a
problem. Here, unless a counterpoise is employed, the connection to ground can be
through a rather long loop; through the coaxial feed cable and the battery supply
leads to the ship's batteries, which are usually grounded to the engine or to a
ground plate connecting to the water.
Fig. 15.55 shows the current return paths for an electrically short antenna on an
automobile. It is clear that only a part of the base current is due to directly induced
currents in the car; a part of the return current flows in the ground beneath the
automobile and returns to the base of the antenna through the capacitance of the
vehicle to ground. The effective ground plane is therefore larger than the physical
size of the frame of the vehicle.

^mrm77Jm///^^

Fig. 15.55 Return path for currents flowing in the base of a short monopole (whip) installed
on an automobile

Fig. 15.56 illustrates the installation of a counterpoise beneath a whip mounted


on a fibreglass or wooden ship. Spilsbury89 has deduced, from an experi-
mental study, that the size of the ground plane must be at least 0.06A in diameter
to be effective, as judged from tuning and matching the antenna. This is a rather
insignificant 'counterpoise'. Awadalla and Maclean5 have studied the input impe-
dance and radiation patterns for a monopole antenna at the centre of an elevated
ground-plane of finite size. They show marked effects when the diameter of this
ground-plane become less than A/2. A part of the observed effect is certainly due
to current flow on the feeder cable (Tilston and Secord95) which must be adequately
choked off. These investigations, however, concern what happens at HF and above.
At MF and below, and especially if the antenna is electrically short, we would
intuitively expect that so long as the capacitance of the counterpoise to the water
ground plane beneath is very much greater than the capacitance of the antenna, then
the counterpoise will be effective. Certainly, since the antenna is only a short distance
above the true ground, the antenna will see an image of itself in this ground, and
the pattern will not be much affected.
VLF, Lf and MF antennas 631

There will certainly, however, be current flow in the wires connecting the
counterpoise to the motor or water ground. This author questions the need for a
counter-poise at all, since the antenna could be fed through a grounded metal
pipe, which itself should be inside PVC plastic pipe since there will be RF current
and voltage on it. In other words, the radiator becomes a monopole with elevated
feed. The feeder cable would be routed up the centre of this metal pipe to decouple
it from the radiator as shown in Fig. 15.57.

Fig. 15.56 Method of 'grounding' for a short monopole antenna (not shown) installed on
the roof of the cabin of a fibreglass ship. The sheath of the coaxial feed is con-
nected to the ground-mat counterpoise. The heavy wires labelled (1) are ground
wires connecting to the ship's ground

15.6.5 Tuning and matching electrically short antennas


Typical whip lengths for land vehicle applications are about 3 m, for operation on
frequencies 2-3 MHz, and for marine applications, whips of length up to 10 m are
typically used for these frequencies. Whips of length 15 m are used for MF mono-
poles for navigational radio beacons and distress calling in the frequency band
405-515 kHz. In this application, top capacitance rods, connected to the base of
the centre loading coil and deflected upwards, are sometimes used to increase the
capacitance of the antenna. The effective capacitance of a 3 m whip is about
30 pF and for the 15 m whip with capacitance rods about 300 pF.
632 VLF, LF and MF antennas

Fig. 15.57 Proposed method of feeding a mo no pole antenna (elevated feed) installed on a
fibreglass ship. The antenna tuning unit (ATU) would be located adjacent to
and connected to the ship's ground

BASE
INSULATOR
TUNE/MATCH I CAPACITY I ANTENNA

T
ll
rrtn

=nnnr>—=L_

rrm
Fig. 15.58 Equivalent circuit illustrating incorrect (a) and correct method (b) of tuning an
electrically short monopole, where a base insulator feedthrough to the inside of
a metal vehicle or an A TU is employed
VLF, LF and MF antennas 633
If the antenna is electrically short and base-tuned, it is absolutely essential that
the tuning coil be installed at the base of the antenna, at the top of the feedthrough
insulator, for metal vehicles. The reason for this is clear from Fig. 15.58. Since the
feed through capacitance can be 3-15 pF, an appreciable non-radiating current
can flow through it if the tuning coil is inside, whereas if the antenna is resonated
'outside' the metal vehicle, the capacitance of the feedthrough insulator becomes
unimportant.
Centre loading circumvents this problem, since the antenna is resonated, and
hence a low impedance. Also, centre loading offers other advantages. Since the
high-voltage part of the antenna is well above the vehicle (see Fig. 15.59), this
could result in reduced losses over base-loading.

Fig. 15.59 Current and voltage distribution on electrically short monopoles that are base-
loaded (left) and centre-loaded (right)

Typical ground loss resistances of about 12 12 at 3 MHz for automobile instal-


lations have been measured. Ground loss resistances of about 2 O are typical for
marine ocean-going vessels.

15.6.6 Ground effects when only a few radial wires are employed
Counterpoises of only a few radials, say less than four, are sometimes used for
transportable applications. The radials are laid on the ground. Fig. 15.12 showed
the effective ground loss resistance for radials from 1 to 120, for a frequency of
about 3 MHz. Counterpoises of three or more radials can be arranged symmetrically
about the electrically short antenna, which exhibits no directional effect. When the
number of radials is reduced to two, the field strength in the plane normal to these
radials is marginally stronger (by 1.5-2 dB at 3.8 MHz, employing radials about
A/4 long) than in the plane containing the radials. When only one radial was used,
its orientation relative to the direction of propagation, had a significant effect on
the transmitted signal strength. When the radial was oriented towards the direction
of propagation, the signal strength was found to be 10 dB higher than when the
radial was directed away from it. Similar directional effects have been observed
for vehicular antennas, where the antenna is placed on one of the 'corners' of the
body of the vehicle.
634 VLF, LF and MF antennas

15.7 Long wire antennas, dipoles and loops


Clearly there are many antenna types which have not so far been discussed in this
review, nor is there space to do so. This Section considers a miscellaneous selection
of three antenna types.

15.7.1 Travelling-wave an tennas


The first long-wire travelling-wave antennas were mounted electrically close to the
earth. The wire was terminated in a resistor equal to its characteristic impedance
at one end, and matched to the receiver at the other end. Antenna lengths of
several wavelengths were employed which provided gain and directivity in the
direction away from the receiver in the direction of the wire antenna. These antennas
for many years were thought to be effective only as receiving antennas (Beverage
et al,18), since their efficiency was so low (< 1%). In more recent years, the wave
antenna has been successfully employed at HF and below as a transmitting antenna
(Seeley and Smith;78 Seeley and Moison;80 and Iitva and Rook 59 ), At HF very
directional arrays have been employed, feeding up to 32 Beverage elements as a
linear-phased array. Not only does directivity gain increase with increase in the
number of elements, but the radiation efficiency also increases. Very directive
antenna arrays can be built, which are easy to construct and are relatively inexpen-
sive. At MF and below, however, the amount of land required for the installation
rapidly becomes prohibitively large with decrease in operating frequency.
Litva and Rook59 have calculated the theoretical patterns (see Fig. 15.60) for
a 125 kHz Beverage antenna situated over poor ground with the following dimen-
sions: height 7.62 m, length 2.4 km. That is, the antenna is one wavelength long.
It is to be noted that the gain of the antenna in the maximum of the lobe is
— 15 dBi. While this is a rather inefficient antenna, the cost of a Beverage antenna
is small compared with a LF tower, and a desirable feature of the Beverage is that
it is broadbanded. The power gain can be increased by feeding more than one
element, as a linear phased array. The expected gain increases by 3 dB each time
the number of elements is doubled. While efficient antennas must be separated
by about A/2 to prevent interaction of their induction fields, inefficient antennas
can be brought much closer together; in fact, studies have shown that interaction
of individual elements does not become appreciable until the elements are spaced
closer than their heights above ground (7.62 m for the 125 kHz Beverage discussed
above). Thus, 4 (or more) Beverage elements operated in phase could be a practical
antenna for some applications. The expected gain is — 11 dBi (or greater if more
elements are employed) for the antenna discussed above.

15.7.2 Inverted- V dipole antenna


The horizontal half-wave dipole antenna radiates a vertically polarised field in the
direction containing the dipole (see Section 15.6.2). A single tower canrbe used
to support an inverted-V resonant dipole antenna. Vertical radiation patterns
measured in planes broadside to, and in the plane of, the antenna are shown in
Fig. 15.61.
VLF, LF and MF antennas 635

80° 60° 40°

20°

200° 340°

240° 260° 280° 300° 320°

160*

180°

200°- 340°

320°

Fig. 15.60 Theoretical vertical and azimuthaf radiation patterns for a Beverage antenna at
125 kHz over poor (dry) soil. The length of the antenna is 2.4 km and its height
is 7.62 m
636 VLF, LF and MF antennas

15.7.3 Half-delta loop antenna


A delta loop antenna used on its fundamental frequency has a perimeter of one
wavelength. A grounded version of this antenna is illustrated in Fig. 15.62. Here
h-\-2h~ A/2, and the antenna including its image in the ground is one wavelength
resonant. This antenna radiates dominately a vertically polarised signal at its
fundamental frequency / 0 . The vertical plane patterns, measured on an antenna
model range, are shown in Fig. 15.63a, measured in the plane of the antenna. The
azimuthal patterns are shown in Fig. 15.63Z? (the plane of the antenna is in the
0-180° axis). In the plane broadside to the delta loop, it radiates approximately
like two A/4 antennas in phase, giving gain in this direction, about 6 dB over a
dipole in free space. Like all loop antennas, it is resonant at / 0 , 2/ 0 , 3/ 0 , 4/ 0 , etc.
The patterns up to 4/ 0 are also shown. The input impedance of the grounded half-
delta loop at/o is about 80 £2, providing a good match to the coaxial cable feeding it.

120 100 80 60 40 20 20 40 60 80 100 120

100 90" 80'


110° 70°
120°
y/ ^ \
T"'
\
y

X
130° ' \

140
X
/ L* 90° — J\ r
127°-"
180 v

kW'.V
150 I
r

160'/
i\
70'/ /

0 i
120
i...
100 80
j i_
60 ;\N
40 20 20 40 60 80 100 120

Fig. 15.61 Elevation patterns (measured on an antenna pattern range) for inverted-V dipole
antennas: (a) horizontal polarisation in the plane broadside to the antenna; and
(b) vertical polarisation in the plane of the antenna

15.8 Receiving antennas

Most of the antennas so far described were discussed in terms of radiation efficiencies
which are important for transmitting application. For receiving the requirements
VLF, LF and MF antennas 637
are quite different. Here signal/noise is to be maximised, where the noise is atmos-
pheric, man-made noise and receiver set noise. In the frequency band of interest,
MF and below, the atmospheric-noise field strengths increase with a decrease in
frequency. At VLF, since atmospheric noise levels are relatively high, there is no
need to have a receiving antenna that is physically larger than the optimum dimen-
sions which will provide a signal that is well above the atmospheric and receiver set
noise levels. Although directivity could help, directivity at MF and below is not
normally used for receiver application, other than the figure-of-eight directivity of a
small loop antenna.

Fig. 15.62 A half-delta loop and its image in the ground

Other considerations are the sensitivity of the antenna to man-made noise, which
in the MF band and below is dominantly electrostatic in nature, and the sensitivity
of the antenna to precipitation and snow static. For aircraft applications, precipi-
tation static can be the major limitation.
A small 1 m square loop is considered to be a very desirable antenna (Belrose
et a/.14), since it can be electrostatically screened, and so provide good immunity
to electrostatic noise fields and precipitation static. For point-to-point fixed fre-
quency operation, a remote preamplifier can be employed, designed to match the
tuned loop to a coaxial transmission line. Hence the loop can be considerably
remote from the main receiving equipment. For portable application, a ferrite loop
provides a considerable reduction in physical size for the same sensitivity (Belrose10).
638 VLF, LF and MF antennas

120'

30°

180°
10 20 30 40 50 60 70 80 90 100

(a)

(b)

Fig. 15.63 Radiation patterns (measured on an antenna pattern range) for a half-delta loop:
(a) Elevation patterns for vertical polarisation in the plane of the delta loop;
(b) Azimuthal patterns for vertical polarisation (the delta loop is in the 0-180°
axis direction}. The various curves are for the second, third and fourth harmonics
of the fundamental frequency f0 of the antenna
VLFf LF and MF antennas 639
Very short rod-type active antennas offer an attractive alternative to loops. The
main advantage of the active monopole antenna is that it is very broadband, but,
being broadband, the principal disadvantage is that of intermodulation in the
preamplifier, due to the strong signals in the broad receiving bandwidth.
A great deal of research has been carried out, for example, by Lindenmeier and
Landstorfer,58 Lindenmeier57 and Nordholt and von Willigen,67 to develop short
active receiving antennas that are extremely linear for high incident field strengths,
so as to reduce the problem of intermodulation.
If the active antenna of height ha is mounted on a short grounded metal mast
of height hm, the effective height, which determines signal pick-up, is greatly
increased. For a short active rod antenna, its effective height, when hm = 0, is

Kff (hm = K
If mounted on a mast

hm +
h
eff 2
(hm, = 0 ) KI2
If ha = 1 m,and hm = 2 m,
heff
= 5 (14 dB)
h
eff >=0)
The present state-of-the-art is such that, according to Lindenmeier and Land-
storfer,58 the basic sensitivities for a 21 cm antenna at 1000, 100 and 10 kHz
are about 22, 45 and 150 nV/m\/Hz. Therefore, in order to obtain a 10 dB SNR
in a receiver bandwidth of 3 kHz, the signal strength at 100 kHz must be 10 juV/m
for the 21 cm active rod antenna. For a i m antenna on a 2 m mast, this minimum
field strength is 0.4 /xV/m.
Unfortunately, the maximum tolerable field strength Et decreases in proportion
with the increase in effective height. That is, for a 21 cm active antenna (hm = 0 )
the maximum field strength ^ f = 55 V/m, but for the 1 m antenna on a 2 m mast,
Et = 2.3 V/m. While this is still a fairly high field strength, it must be remembered
that, in actual fact, the rms field strength of all strong signals in the bandwidth
of the active antenna may exceed this level, which is the reason why active antennas
must be used with due caution. Taking nonlinear effects such as intermodulation
and cross-modulation into account, present research has been directed toward the
development of a tunable short active receiving antenna. Hopf and Lindenmeier41
have presented design considerations for an extremely linear selective active antenna
with at least two tuned resonant circuits, tunable within the range 1-30 MHz
within milliseconds. For a 1.2 m active antenna, the second-order intermodulation
at the receive frequency fr is typically — 70 dB if undesired signals at 0.45/ r and
O.55/r, with field strengths of 100V/m each, excite the antenna. Considering
640 VLF, LF and MF antennas

third-order intermodulation distortion produced by undesired signals with fre-


quency differences of 10 and 20% from the receive frequency and 100 V/m, this
distortion is — 65 to ~ 75 dB.

15.9 Practical antenna design

The full potential of VLF and LF transmitting stations can only be attained if the
radiating system is designed to give maximum efficiency. The degree to which this
can be attained, however, is a function of many sometimes conflicting factors such
as land availability, permissible tower heights, bandwidth required, the power of
the transmitter and the capital outlay. The compromises necessary have to be
considered and all factors optimised. The electrical and mechanical structural
design of the antenna, once the antenna type and the site for it is chosen, normally
requires the combined effort of radio, civil and mechanical engineers. The practical
design of such antennas is a specialised field of engineering practised by a few
major companies that supply complete systems. If the antenna design has been
done, companies supplying towers can erect them and string the antenna wire. MF
antenna design is also a field of specialisation, particularly if, as already discussed, a
complicated directional array is required to meet licensing requirements. In this
case, the design and erection contracting, the field testing and the adjustment of the
array, is normally done by broadcast consultants specialising in this particular field.
Obviously, we cannot go into much detail here concerning practical antenna
design, except to make a few general remarks, and illustrate photographically some
of the elements of antennas, as well as of complete systems.

15.9.1 Guy ed radio an tenna mas ts


Antenna masts of height from 30 to 450 m are employed. While some of the earlier
installations, and even today particular installations, employed self-supporting radio
towers and masts, most installations employ guyed masts, especially if they are to
be base-insulated. For critical design, the current is more nearly sinusoidal on a
guyed tower of uniform cross-section. The lattice towers are usually of square
cross-section. For a 100 m mast, a typical face width is about 1.5 m, although
masts which support a particularly heavy load would have a greater cross-section
and heavier construction.
A few masts are cylindrical, but this is the exception. The OMEGA mast at
Tsushima, Japan, is made up of 68 cylindrical sections 3 m in diameter, each weigh-
ing from 17.3 x 103 kg for the lower sections to 14.5 x 103 kg for the upper
sections. It is guyed at six levels, and has a height of 450 m.

15.9.2 Base insulators for guyed masts


Base insulators are made of high-strength porcelain and some are oil-filled to
eliminate RF loss due to water condensation on the inner surfaces. The outer
surface is polished smooth, to minimize leakage and flashover. Various types are
illustrated in Fig. 15.64. Those made by some companies are of cylindrical shape,
VLF, LF and MF antennas 641

Fig. 15.64a A pair of conical base insulators with spark gap for lightning protection

others are a curved-sided cone and some have corrugated surfaces. High-voltage
and high-power installations are fitted with corona rings and most have rain shields.
The oil-filled types can be fitted with heaters. Base insulators for VLF and LF
antenna installations are manufactured to provide 9 x 103 to 4.5 x 106 kg working
loads for a safety factor of about 3. They are usually provided with a lightning
spark gap and an isolation transformer for tower lighting. Design working voltages
range from about 25 kV to 500 kV.
A typical installation for transmitting powers of 100 kW is shown in Fig. 15.65.
Note the oversized rain shield and lightning spark gap.
Is is clear that, especially for the larger installations, an antenna design which
utilises a grounded tower offers distinct advantages since no base insulator, lightn-
ing spark gap or lighting isolation transformer are required. Clearly, cages about
towers are needed, but strain insulators can be used. The mast can also be used to
support VHF and UHF antennas.
642 VLF, LF and MF antennas

Fig. 15.64b Base insulator with rain shield

<*tm#tffr' n

Fig. 15.64c Base insulator with corona rings


VLF, LF and MF antennas 643

Fig. 15.64d Base insulator with rain shield and spark gap for lightning protection

15.9.3 Guy strain insulators


Most guy insulators are of simple design. They comprise a ceramic or procelain
rod fitted with corona rings and rain shields, as shown in Fig. 15.66a. The heavier-
duty type are oil-filled. Typical working loads range from 1.8 to 125 x 103 kg.
Recommended operation voltages, wet, range from 50 kV to 100 kV. For high
working voltages, two or more guy insulators would be operated in series. These
are the insulators that are mounted next to the mast.
The guys are broken up with fail-safe-type insulators, either the basic johnny-
ball type (Fig. 15.665) or compression cone type for heavier working loads (Fig.
15.66c).
For smaller installations, glass-fibre insulators or epoxy-type guys (phillistrand)
provide working strengths up to 8.3 x 103 kg.
In the case of major installations at VLF or LF where insulated guy sections are
long, porcelain-housed static drain insulators are sometimes required. These are
jumpered across the guy insulators as shown in Fig. 15.67. These static drain
resistors prevent flashover due to static charge build-up on the insulated guy sec-
tions by the action of wind, rain or snow. Units having resistances from 75 kO to
5 MO are used depending on operating voltages.
644 VLF, LF and MF antennas

Fig. 15i>4e Base insulator with a ring type isolation transformer for tower lighting

15.9.4 Top-loading design


A top-hat suspended between towers is insulated from the towers by strain insu-
lators of the type described above. The mechanical loading of members of VLF
and LF top-hats is rather large, and it becomes necessary to use high-strength
conductors, such as stranded phosphor bronze or copper-clad steel where excep-
tional strength is required.
The proper tensioning of the top-hat is an important aspect of the mechanical
design of the antenna. If the down lead is not a central tower but a cage of wires,
the top-hat can be rigidly connected, by means of strain insulators, to the grounded
towers, but with some slack. The cage down lead is held under tension by a counter-
VLF, LF and MF antennas 645
weight at its lower end, below a series strain insulator which is effectively the 'base
insulator'. The top-hat for major installations is not rigidly connected to the sup-
porting towers, but to cables which run over pulleys at the top of the mast,

Fig. 15.65 Photograph of the base of a 100 kW LF umbrella type radiator (after Marconi
Communication Systems, Ltd)

and down the centre of the mast, or alongside it where they are attached to counter-
weights to optimise the loading and tension. This arrangement protects the top-
loading, since, under extreme icing conditions, the top-hat will be automatically
lowered. In major VLF installations, these counterweights can be tremendous, up
to 180 x 103 kg. The strain insulators holding the flat top are equally impressive,
646 VLF, LF and MF antennas

F ig. 15 J66a Oil- filled sa fety co re type of guy strain insulato r


s
I

F ig. 15.66b Johnn y-ball types of guy strain insulator Fig. 15.66c Compression-type cone guy strain insulator
648 VLF, LF and MF antennas

comprising 16 fail-safe compression-type insulators in a string, fitted with corona


rings, weighing about 6.8 x 103 kg per string (see Fig. 15.68).

Fig. 15.67 Static-drain resistors for guyed radio masts (after Lapp Insulator Division, Inter-
pace Corporation)

The configuration of the top-hat wires also requires some ingenuity in the case
of major installations. The arrangement is sometimes configured so that, by use of
strain insulators, it is broken up into insulated loops, each loop having its own
down-leads to provide sleet melting circuits.

15.9.5 Antenna tuning units


It is customary for the output from the transmitting station to feed the radiating
system by means of a solid or an air-insulated coaxial cable. This arrangement
enables the relative siting of the antenna and the transmitter building to be opti-
mum with regard to access roads, power supplies and other services.
It is necessary, therefore, to tune the antenna and match it to the impedance of
the coaxial feeder cable. While various methods are employed, the most common
VLF, LF and MF antennas 649

Fig. 15.68 Giant antenna insulator for top-loading, manufactured for USN VLF transmitter,
Cutler, Maine (after Lapp Insulator Division, Interpace Corporation)

MATCH TUNE I ANTENNA

TUNE/MATCH ANTENNA

b) /TT77
Fig. 15.69 Typical method of tuning and matching VLF/LF antennas: (a) Equivalent circuit
of antenna, base tuning coil and L-section matching unit; (b) actual components
required
650 VLF, LF and MF antennas

Fig. 15.70a Helix tuning coil for 100kW LF radiator (after Marconi Communication Systems
Ltd)

type, if the antenna is capacitively reactive, is an L-section match, which provides


a shunt coil to ground so the antenna is operated at DC ground. The circuit arrange-
ment from the point of view of visualising and calculating the circuit parameters
for tuning and matching is shown in Fig. 15.69a. Here, the coil Lo resonates the
antenna, and the Irsection matches the resistive impedance Rt = Ra + Rc to the
impedance of the coaxial feeder cable Ro. In actual fact, the series arm of this
L-match section (C m ) is not needed since the antenna can be tuned to provide the
required capacitive reactance (Fig. 15.69Z?). In the actual circuit

Xx = Xa-XCm = Xa-y/Rt(R0-Rt)

and X2 - XLm - Ro
Ro —Rt
VLF, LF and MF antennas 651

Fig. 15.70b Helix tuning coil for 2000kW VLF radiator (after Continental Electronics
Manufacturing Co.)

The main tuning inductor is usually wound with litz wire. Typical installations
employ antenna tuning coils having diameters from about 1 to 6 m, and employ
litz wire having diameters from about 1 to 9 cm. Fig. 15.70 shows photographs
of 100 kW and 2000 kW helix coils.
To facilitate tuning, a variometer coil is generally inserted in series with the
tuning coil, on the low-voltage side (the end towards the transmitter). Fig. 15.71
shows variometer coils for the power levels noted above.
The shunt inductor is usually wound with copper tubing to facilitate tapping
and matching. It is sometimes silverplated. For power levels up to about 25 kW,
a ferrite transformer is sometimes used to match Ro and Rt, once the antenna has
been resonated.

15.9.6 Arrangement of the antenna tuning unit


A typical layout is shown in Fig. 15.72 (Jinkings,45). Variations will occur depend-
ing upon the type of antenna used, the frequency and the power level. In many
installations, the entrance insulator (Fig. 15.73) is through the side of the tuning
hut, often through a heavy glass plate to reduce its capitance and increase the
operating voltage. A Faraday shield is sometimes used instead of copper sheeting.
These are vertical aluminum rods mounted on insulators around the inside walls
652 VLF, LF and MF antennas

Fig. 15.71a Variometer for a 100 kW LF installation


of the tuning hut. They are joined at their lower ends, but not by a complete loop
around the coil. The Faraday screen is divided into eight sections, and each section
is joined at the bottom ends of the rods to a heavy copper strap which runs diago-
nally to the centre of the main tuning helix, where a single wider strap joins this
junction point to the radial ground system.
The tuning hut can be made of dry wood, but certainly not plywood, since
induction heating by the RF could set it on fire. If the entrance insulator enters
through the side of the tuning hut, a wooden or plastic canopy should be used
as a rain shield.

15.9.7 Feeding a single tower radiator at two MF frequencies


MF broadcast stations have power levels from 250 W to 2 MW, but 50 kW is typical
of 'high power' installations in North America. The antennas employed range from
VL F, LF and MF an tennas 653

Fig. 15.71b Variometer for a 2000kW VLF installation (part of the antenna tuning system
illustrated in Fig. 15.70b)

a simple quarter-wave vertical, to multi-tower arrays of up to 8 towers phased


to provide a certain specified pattern. Clear channel stations employ a 5X/8
radiator or a sectionalised monopole up to X in height, so-called anti-fade
antennas.
Since the cost of erecting a 5X/8 mast for use near the bottom end of the broad-
cast band is considerable, transmitters are sometimes diplexed into the same mast
(cf. Laport,56). The circuit diagram in Fig. 15.74 shows the diplexer used by
CBC MF-AM broadcasting stations CBL/CJBC. The transmitters operate on 740
and 860 kHz. The basic design, as one can see, is fairly straightforward. Each
arm has shunt-to-ground circuitry that provides a low impedance (series tuned)
for the unwanted frequency and a high impedance (the reactance of the series
tuned arm is parallel-tuned) for the wanted frequency. Similarly, the series arm
has series-tuned elements for the wanted frequency, and this combination is
654 VLF, LF and MF antennas
TO ANTENNA

© 26 SWG COPPER
SHEET CLADDING

FEEDER ''
POWER SUPPLIES
IN
LOADING PAD

Fig. 15.72 A typical LF antenna tuning installation shown with the components inside
the tuning hut (after Jink ings,45)

Fig. 15.73 Antenna feedthrough or entrance insulators (after Lapp Insulator Division,
Interface Corporation)
VLF, LF and MF antennas 655

parallel-tuned to the unwanted frequency. An isolation between the modulated


transmitters of about 50 dB is achieved, with very low coupling loss since silver-
plated copper inductors and vacuum capacitors are used.

Zaf1 1S8H25
Zaf2 56-M37

Fig. 15.74 Circuit for a MF dip/exer-coupler (after CBC, 1974)

15.9.8 Unique installations


There are many examples of exceptional installations, 'unique' being in ~t in
the view of the eyes of the beholder. To conclude this Chapter, two such instal-
lations will be briefly described.
The NATO VLF station near Narvik, Norway, is located within a mountain
valley open to the ocean. The antenna, see Fig. 15.75, utilises two mountains
as anchor points, providing a span for the flat top of 2134 m, 914 m above the
sea floor.
Fig, 15.76 shows a photograph of a 750 kW 6-element monopole Yagi antenna
at Qatar. The station operates on 954 kHz. The drive element is a folded mono-
pole 72 m high, the reflector is 79.25 m and each director is 67 m high. The spacing
is uniform 79.25 m (O.25X). The antenna has a power gain of 8 dB over a single
tower radiator.
656 VLF, LF and MF antennas

Fig. 15.75 Photograph of a fiord mountain supported VLF antenna, Narvik, Norway (after
Continental Electronics Manufacturing Co.)
VLF, LF and MF antennas 657

Fig. 15.76 Photograph of 6-element MF broadcasting Yagi-type monopole array, Qatar,


(after Marconi Communication Systems)
658 VLF, LF and MF antennas
15.10 References

1 ABBOTT, F. R.: 'Design of optimum buried-conductor RF ground systems*, Proc. IRE,


40, 1952, p. 846
2 AIZENBERG, G. S.,BELOUSOV, S. P.,YAMPOLSKY, V. G.,andLINDERBERG, A. Ch.:
'MF antifading antenna with controlled current distribution' Radiotekhnika, 12,
1961,p.21
3 ALEXANDERSON,E. F.: 'Trans-oceanic radio communications', Proc. IRE, 8, 1920,
p. 263
4 ALFORD, A., and FRENCH, E., 'Some observations concerning the re-radiation of
radio frequency energy from power lines and transmission towers* Andrew Alford
Consulting Engineers, Boston, Mass., 6 Aug. 1966
5 AWADALLA, K. H., and MACLEAN, T. S. M.: 'Monopole antenna at centre of a circular
ground plane: input impedance and radiation pattern', IEEE Trans. Ant. & Prop.,
AP-27, 1979, p. 151
6 BALLANTINE,S.: 'On the radiation resistance of a simple verticle antenna at wave lengths
below the fundamental', Proc. IRE, 12,1924, p. 823
7 BALLANTINE, S.: 'On the optimum transmitting wave length for a vertical antenna over
perfect earth', Proc. IRE, 12,1924, p. 833
8 BELOUSOV, S. P., and KOUZNETSOV, V. D.: 'Wideband antifade MF antenna', IEE
Conf. Publication 169, Part 1,1978, p. 379
9 BELROSE, J. S.: 'Short antennas for mobile operation', QST, Sept. 195 3, p. 30
10 BELROSE, J. S.: 'Ferromagnetic loop aerials', Wireless Engineer, 32, 1955, p. 41
11 BELROSE, J. S.: 'Low and very low frequency radio wave propagation', AGARD
Lecture Series XXIX, July 1968, NTIS AD676788, Springfield, VA
12 BELROSE, J. S.: 'Transmission-line low profile antennas', QST, Dec. 1975, p. 19
13 BELROSE, J. S.: 'A kite supported 160- (or 80-) meter antenna', QST, March 1981, p. 40
14 BELROSE, J. S., HATTON,W. L., McKERROW,C. A., and THAIN, R. S.: 'The engi-
neering of communication systems for low radio frequencies', Proc. IRE, Vol. 47,
1959,p.661
15 BELROSE, J. S., LAVRENCH, W., DUNN, J. G., TRUEMAN, C. W., and KUBINA, S. J:
'The effects of re-radiation from high-rise buildings and transmission lines upon the
radiation pattern of MF broadcasting antenna arrays', AGARD Conf. Proceedings
269,1979,p.2-1
16 BELROSE, J.S,, TRUEMAN, CW., KUBINA, S. L, LAVRENCH, W.} and GUNN, J. G.:
'On minimizing the effects of re-radiation from power transmission lines on the radi-
ation pattern of MF-AM broadcasting antenna arrays', IEE Conf. Proc. 195, 1981,
p. 33
17 BEMSD. J.: 'A cage antenna with controlled radiation patterns in the vertical plane'
CCIR Documents 1970-1973, Doc. 10/62, 1972
18 BEVERAGE, H. H., RICE, C W., and KELLOGG, E. W.: 'The wave antenna: a new type
of highly directive antenna', AIEE Trans. 42, 1923, p. 215
19 BOYER, J.M.: 'Hula-hoop antennas: A coming trend?', Electronics, 11, Jan. 1963,
p. 44
20 BROWN, G. H.: 'A critical study of the characteristics of broadcast antennas as affected
by antenna current distribution', Proc. IRE, 24, 1936, p. 48
21 BROWN, G. H.: 'Directional antennas',Proc. IRE, 25, 1937, p. 78
22 BROWN, G. H., LEWIS, R. F., and EPSTEIN, J.: 'Ground systems as a factor in antenna
efficiency', Proc. IRE, 25, 1937, p. 753
23 BROWN, H., and WOODWARD, O. M.: 'Experimentally determined impedance charac-
teristics of cylindrical antennas', Proc. IRE, 33, 1945, p. 257
VLF, LF and MF antennas 659
24 BURKE, G. J.: The numerical electromagnetic code (NEC)', in 'Application of method of
moments' (ed. STRAIT, B. J.), The SCEE (Southeastern Centre for Electrical Engi-
neering Education) Press, St. Cloud, Florida, 1980
25 BURKE, G. J., and POGGION, A. J.: 'Numerical electromagnetic code (NEC) - Method
of moments', Techn. Doc. 116, prepared for the Naval Electronic Systems Command
(ELEX 3041), 1980
26 BURKE, G.J., MILLER, E. K., BRITTINGHAM, J. N., LAGER, D. L., LYTLE, R. J.,
and OKADA,J. T.: 'Antennas near ground', Tech. Report UCID-18626, Lawrence
Livermore Laboratory, 1980
27 BURTON, R. W., and KING, R. W. P.: 'Theoretical considerations and experimental
results for hula-hoop antenna',Microwave J., Nov. 1963, p. 89
28 CARTER, P. S.: 'Circuit relations in radiating systems and applications to antenna prob-
lems', Proc. IRE, 20, 1932, p. 1004
29 COX,C. R.: 'Mutual impedance between vertical antennas of unequal heights', Proc.
IRE, 35, 1947, p. 1367
30 EVANS, W. M.: 'A transportable LF communications system'Communications Research
Centre, Ottawa, CRC Report 1213, 1971
31 MA, M. T. and FITZGERRELL, R. G.: 'Design of a van-top low-profile HF antenna' US
Dept. of Commerce, Office of Telecommunications Report 77-131, Oct. 1977
32 FLOOD, D. P., and FIELD, J.C.: 'Experimental study of the backscattering from con-
ducting objects in the range of one-quarter to one wavelength' AFCR Report 64-541,
April 1964
33 FOURNIER, M., and POMERLEAU, A.: 'Experimental study of inductively loaded short
monopole',IEEE Trans., VT-27, 1978, p. 1
34 FRANCHI, P. R.: 'Air Force VLF communication antennas', Proc. ECOM-ARO Workshop
on electrically small antennas (ed. G. GOUBAU and F. SCHWERING), US Army Elec.
Command, Fort Monmouth, NJ, 6-7 May 1976, p. 90
35 GANGI, A. F., SENSIPER, S., and DUNN, G. R.: 'The characteristics of electrically short
umbrella top-loaded antennas', IEEE Trans., AP-13, 1965, p. 864
36 HALLEN,E.: 'Theoretical investigations into transmitting and receiving qualities of
antennae',Nova Ada RegiaeSoc. Sci. Upsaliensis, Ser. IV, 11, 1938, p. 1
37 HANSEN,R.C: 'Optimum inductive loading of short whip antennas', IEEE Trans.
VT-24, 1975, p. 21
38 HARRINGTON, R. F.: 'Field computation by moment methods' Macmillan Co., NY,
and Collier-Macmillan, Ltd., London, 1968
39 HATCH, J. F., STRUSZYNSKI, W., and THURGOOD,H.: 'The Marconi eight aerial
Adcock HF direction finder type S4&0\ Marconi Rev. XXIX, 1966, p. 1
40 HEAD,H.T.: 'Medium-frequency broadcast antennas', in 'Antenna Engineering Hand-
book' (ed. H. JASIK) McGraw Hill Book Co., 1961, p. 20-1
41 HOPF,J.F., and LINDENMEIER, H. K.: 'Extremely linear electrically tuneable active
receiving antenna'. IEE Conf. Pub. 195,1981, p. 75
42 HOWE, G.W. O.: 'On the capacity of radio-telegraphic antennae', Electrician, 73, 1914,
p. 829
43 JAGGARD, D. L.: 'An application of isoperimetric inequalities to the calculation of
equivalent radii' National Radio Science Meeting, Boulder, Colorado, 1979, p. 198
44 JEFFERS,C. L.: 'An antenna for controlling the nonfading range of broadcast signals',
Proc. IRE, 36, 1948, p. 1426
45 JINKINGS, P. W.: 'Antenna systems for LF transmitters', Point-to-Point {Marconi Rev.),
Jan. 1971,p.32
46 JORDAN, E. C, and BALMAIN, K. G.: 'Electromagnetic waves and radiating systems',
Prentice-Hall, Inc. 1968
47 KANDOIAN, A. G., and SICHAK, W.: 'Wide-frequency-range tuned helical antennas
and circuits' IRE Nat. Conv. Record, 1953, p. 42
660 VLF, LF and MF antennas

48 KING, R. W. P.: 'Measured admittances of electrically thick monopoles', IEEE trans.


AP-20,1972, p. 763
49 KING, R. W. P., and HARRISON, C. W.: 'Antennas and waves: A modern approach'
MIT Press, 1969
50 KNIGHT, P.: "The design of cage-driven MF aerials', Electronic Eng., Feb. 1966, p. 82
51 KNIGHT, P.: "The design and adjustment of MF broadcasting aerials', Electronic Eng.,
Jan. 1967, p. 4
52 KOSSEY, P. A., LEWIS, E, A. and FIELD, E. C : 'Relative characteristics of TE/TM waves
excited by airborne VLF/LF transmitters', AGARD Conf. Proceedings 305, 1982,
p. 19-1
53 LABUS,J.: 'Mathematical calculation of the impedance of antennas', Hochf. und Elek,
41, 1933, p. 17
54 LACHARNAY, S.: 'New high-gain MF antenna for Lille-Camphin broadcasting transmitter
(a quarter-wave skirt antenna)', Rev. Fr. Radio Diff. et Telev., (11), 1969
55 LACHARNAY, S.: 'Modern antennae for LF, MF and HF waves', Rundfuntechn. Mitteil,
13, 1969, p. 40
56 LAPORT, E. A.: 'Radio Antenna Engineering', McGraw-Hill Book Co., 1952
57 LINDENMEIER, H. K.: 'The short active receiving antenna an appropriate element for
application in antenna arrays' IEE Conf. Proc. 169, 1978, p. 91
58 LINDENMEIER, H. K., and LANDSTORFER, F. M.: 'Some examples of small, low-
noise, highly linear active antennas produced in quantities for various applications',
Proc. ECOM-ARO Workshop on electrically small antennas (ed. G. GOUBAU and
F. SCHWERING) US Army Electronics Command, Ft. Monmouth, 1976, p. 105
59 LITVA,J., and ROOK, B. J.: 'Beverage antennas for HF communications, direction
finding and over-the-horizon radars' Communications Research Centre, Ottawa, CRC
Report 1282, 1976
60 LO, Y. T.: *A note on the cylindrical antenna of noncircular cross section', /. Appl
Phys. 24,1953, p. 1338
61 MARTIN, C. A,, and CARTER, P. S,: 'Low frequency antennas', in 'Antenna Engineering
Handbook' (ed. H. JASIK). McGraw4M Book Co., 1961
62 MARTIN, L., CASTILLO, J. P., and LEE, K, S, H.: 'Broad-band analysis of VLF/LF
aircraft wire antennas', IEEE Trans., AP-26,1978, p. 141
63 MONTEATH,G. D.: 'Application of the compensation theorem to certain radiation and
propagation problems', Proc. IEE, 98 Pt. IV, 1951, p. 23
64 MORRISON, J. F., and SMITH, P. H.: 'The shunt-excited antenna', Proc, IRE, 25, 1937,
p. 673
65 NIVER, E., and HIZAL, A.: 'The effects of stratified ground on characteristics of inverted
L antenna' AGARD Conf. Proc. 269,1979, p. 1-1
66 NOLAN, S. U.: 'Developments in MF radiator design', Sound and Vision Broadcasting, 15,
1974, p. 1
67 NORDHOLT, E. H., and VON WILLIGEN, D.: 'A new approach to active antenna design',
IEEE Tram. AP-28, 1980, p. 904
68 PAGE, H., and MONTEATH, G. D.: 'The vertical patterns of medium wave broadcasting
aerials',Proc. IEE, B102,1955, p. 279.
69 PRESTHOLD, O. L.: 'The effects of re-radiation from television towers and other struc-
tures upon directional antennas' Proc. NAB broadcast Engineering Conference, 1958
70 RAGHURAM,R., SMITH, R. L., and BELL,T. F.: 'VLF Antarctic antenna: Impedance
and efficiency',IEEE Trans. AP-22, 1974, p. 334
71 RAMSDALE,P. A.: 'Wire antennas', in Advances in Electronics and Electron Physics,
Academic Press, Inc., 1978, p. 123
72 RAY, H. A.: 'Dispensible transmitting antenna VLF/LF investigation' Rome Air Develop-
ment Centre, Griffiss Air Force Base, New York, Final Report RADC-TR-72-28, Mar.
1972
VLF, LF and MF antennas 661
73 RICHMOND, J. H.: 'Radiation and scattering by thin-wire structures in the complex
frequency domain', NASA Contractor Report CR-2396,1974
74 RICHMOND, J. H.: 'Computer program for thin wire structures in a homogeneous con-
ducting medium', NASA Contractor Report CR-2399,1974
75 RICHMOND, J. H.: 'On the edge mode in the theory of thick cylindrical monopole
antennas',IEEE Trans, AP-28, 1980, p. 916
76 ROYER, G. M.: 'The effects of re-radiation from high-rise buildings and towers upon the
antenna patterns for an AM broadcast array*, .1981 IEEE Int. Symp. on Electro-
magnetic Compatibility, Boulder, Col., Aug. 1981
77 SATHER, O. J.: 'Construction of the WOR transmitter directional array', IEEE Trans.
BC-15, 1969, p. 65
78 SEELEY, E. W., and SMITH, R. D.: 'Radiation efficiency of variable-wave velocity hori-
zontal antenna near the earth', Naval Ordnance Laboratory, Corona, Calif., Report 730,
AD656397,1967
79 SEELEY, E.W., and WIBORG,P. H.: 'Horizontal VLF transmitting antennas near the
earth' Naval Ordnance Laboratory, Corona, Calif.,NOLC Report 721, 1967
80 SEELEY, E. W., and MOISION, W. K.: 'Horizontal end-loaded VLF transmitting antenna',
Naval Weapons Laboratories, Tech. Pub. 782, AD839107, 1968
81 SCHELKUNOFF, S. A.: 'Theory of antennas of arbitrary size and shape', Proc. IRE,
29, 1941, p. 493
82 SIEGEL, E., and LABUS,J.: 'Apparent resistance of antennas', Hochf. und Elek., 43,
1934, p. 166
83 SIEGEL, E., and LABUS, J.: 'Transmitting Antennas', Hochf. und Elek., 49, 1937, p. 87
84 SMEBY, L. C : 'Short antenna characteristics- Theoretical',Proc. IRE, 37, 1949, p. 1185
85 SMITH, C. E.: 'Theory and design of directional antennas' Smith Electronics Inc., 1969
86 SMITH, C. E.: 'Design and operation of directional antennas'. Smith Electronics, Inc.,
Cleveland, Ohio, August, 1969 (Selected articles on directional antennas)
87 SMITH, C. E., and GRAF, E. R.: 'Increased capacity for VLF umbrella antennas using
multi-wire rib construction',IEEE Trans., AP-16, 1968, p. 766
88 SMITH, C. E., and JOHNSON, E, M.: 'Performance of short antennas', Proc. IRE, 35,
1947,p.1026
89 SPILSBURY, A. J.: 'Ground systems for high frequency radio telephones on wood and
fibreglass ships'. Spilsbury Communications, Vancouver, 30 Aug. 1979
90 SURUTKA, J. V. and POPOVIC, B, D,: 'Field intensity versus radiated power, height
and thickness of a linear antenna', Proc. IEE, 114,1967, p. 55 3
91 SWANSON, E. R., BRITT,J.E., and SMITH, A. N.: 'Omega possibilities: Limitations,
options and opportunities', NOSC Tech. Report 283, Naval Ocean Systems Centre,
San Diego, June 1978
92 TAJ,C. T.: 'Characteristics of linear antenna elements', in 'Antenna Engineering Hand-
book' (ed. H. JASIK), McGraw-Hill Book Co., 1961, p. 3-1
93 TENG,C, and KING, R. J.: 'Surface fields and radiation patterns of a vertical electric
dipole, over a radial ground system', AGARD Conf. Proc. No. 269, Terrain profiles
and contours in electromagnetic wave propagation. Tech. Editing and Reproductions
Ltd., 1979, p. 4-1
94 THIELE,G. A., EKELMAN, E. P., and HENDERSON, L. W.: 'On the accuracy of the
transmission model of the folded dipole', IEEE Trans. AP-28, 1980, p. 700
95 TILSTON, W. V., and SECORD, A. H.: 'The radiation patterns of ground rod antennas',
Electron, and Commun., Aug. 1967, p. 27
96 TIPPE, VonW.: 'Zur Dimensionierung fur Verticale LW-BZW. MW-Monopolantennen',
Rudfunktech. Mitleil 24, 1980,p.154
97 TRUEMAN,C. W., and KUBINA, S. J.: 'Prediction by numerical computation of the re-
radiation from and the detuning of power lines', Tech. Report TN-EMC-81-03,
Concordia University, Montreal, 1981
662 VLF, LF and MF antennas

98 TRUEMAN,C.W., KUBINA, S. L, and BELROSE, J. S.: 'The prediction of re-radiation


from power lines at MF frequencies', IEEE Int. Symposium on EMC, Baltimore, Md.,
1980
99 TRUEMAN, C. W., KUBINA, S. J., and BELROSE, J. S.: 'Resonant behaviour and detun-
ing of power lines in the MF band*, IEEE Int. Symp. on Electromagnetic Compatibility,
Boulder, Colorado, Aug. 1981
100 WAIT, J. R., and POPE,W. A.: 'Characteristics of a vertical antenna with a radial con-
ductor ground system' in 'Applied Scientific Research, Vol. B4, Martinus Nijhoff,
1954
101 WAIT, J. R., and CONDA, A. M.: 'Pattern of an antenna on a curved lossy surface',
IRE Trans., 1958, AP-6, p. 348
102 WATT, A. D.: 'VLF engineering', Pergamon Press, 1967
103 WEEKS, W. L., and FENWICK, R. C : 'Submerged antenna performance', IRE Inter-
national Conv. Record 1962, Vol. 10, p. 108
104 WHEELER, H. A.: 'Small antennas',IEEE Trans., AP-23, 1975, p. 462
105 WILLIAMS, H. P.: 'Antenna theory and design Vol. 1 and 2', Sir Isaac Pitman, 1950
106 MONTEATH, G. D.: 'The effects of the ground constants, and of an earth system, on the
performance of a vertical medium-wave aerial' Proc. IEE, 105C, 1958, p. 292
Chapter 16

High frequency antennas


M. F. Radford

16.1 Introduction to HF

16.1.1 HF propagation
The HF band as defined by the CCIR extends from 3 to 30 MHz 100 to 10 m
wavelength, but for practical purposes it is the band for which the ionospheric
skywave is the dominant mode of propagation. The useful extent of the HF band at
any given time depends upon the ionospheric conditions, and the limits may extend
beyond the CCIR decade from below 2 MHz to over 40 MHz. In order to maintain
communication over a given path as the ionosphere varies it is necessary to have
available a choice of frequencies spanning most of the band, and 4 to 28 MHz is a
typical practical system bandwidth requirement.
Since the use of ionospheric propagation has an important bearing on the design
of HF antennas, this chapter will begin with a discussion of the ionosphere. The
ionosphere is a sparse region which extends from around 50 km to around 400 km
in altitude. During daylight hours it is bombarded by solar radiation so that the
rarefied air becomes ionised by the ultra-violet and X-ray emissions. Different parts
of the solar spectrum are absorbed at different altitudes, so that several ionised
layers are formed. The upper layers reflect radio waves while the lower layers
attenuate the waves passing through them (Fig. 16.1).
Reflection can be explained by considering the radio refractive index of an
ionised layer. The refractive index JJL of a plasma of ionisation density TV electrons
per cubic metre is given approximately by

• - ! ? 06.1)
where / is the frequency in hertz. For any given density of ionisation the refractive
index decreases as the frequency decreases, and at a frequency/p called the plasma
frequency, where

/ p = SIN (16.2)
664 High Frequency Antennas

the refractive index (JL falls to zero. A radio wave at a frequency below/ p and norm-
ally incident on an ionised layer penetrates only as far as the point where JJL falls to
zero, and is then totally reflected. The plasma frequency at the peak of the layer is
called the critical frequency/ c . Normally-incident waves at frequencies above/ c
pass right through the layer.
v.h.f. transmitted
reflecting layer

m.f. absorbed h.f. reflected

Fig. 16.1 Reflection from the ionosphere

At frequencies above the critical frequency, reflection may still occur if the wave
is obliquely incident. The phenomenon is similar to that of total internal reflection
of light in a water surface. The highest frequency which can be totally reflected at
a given angle of incidence 0 is given approximately by
/ = /csec0 (16.3)
and may be several times greater than the critical frequency. This formula is only
approximate because of the curvature of the earth.
The attenuation loss depends mainly on the product of the ionisation density
and the frequency of collisions between the free electrons and heavier particles.
When the electrons are set in motion by a radio wave they acquire oscillatory energy
from the wave, but when the collisions occur this energy is converted into heat.
Since the energy comes from the wave, the latter is attenuated. In the HF band the
rate of attenuation A is given approximately by
1.16 x 1(T6 Nv , , , ,
A = j dB/metre (16.4)

where v is the electron collision frequency. The attenuation becomes substantial at


the lower end of the HF band and determines the lower useful limit, just as the
refractive-index behaviour determines the upper useful limit.
The behaviour of the ionosphere is modified by the earth's magnetic field. This
causes the free electrons to follow elliptical paths, and consequently the polarisation
of the wave changes as it progresses through the ionosphere. There are, however,
two elliptical wave polarisations which can travel through the ionosphere unchanged;
they are known as the 'ordinary' and 'extraordinary' waves. Any wave incident on
the ionosphere can be resolved into these two polarisations and the propagation of
the two waves can then be considered independently until they emerge, when they
High Frequency Antennas 665
recombine in a random manner as described in Section 16.1.2. The refractive index
of the ionosphere is slightly different for the two waves and the rate of attenuation
of the extraordinary wave is greater than that of the ordinary wave.
Since the ionisation density of the upper atmosphere can be estimated from the
time, season, latitude and sunspot number, it is possible to forecast critical frequen-
cies and attenuations with sufficient accuracy for practical purposes for most of
the time. However, the ionosphere is subject to violent changes when flares occur
on the sun, due both to the increased electromagnetic radiation and to the bombard-
ment by heavier particles some 36 hours later. These phenomena are relatively
unpredictable, and may lead to a complete loss of HF communication over the
whole band when the solar activity is sufficiently intense.
To determine the actual path of the skywave it is necessary to consider the
altitudes and densities of the various layers in a little more detail. There are three
principal layers which affect HF propagation, the D, E and F layers. There are in
fact no A or B layers, but the early research workers left space in their nomenclature
for subsequent discoveries (Fig. 16.2).

400

300

~ 200

100

0
12 18 24
local time , hrs
Fig. 16.2 Ionospheric layers

The D layer occurs at around 80 km altitude. It is formed in the morning and


decays in the evening. It has a relatively low ionisation density, N being around
10 9 /m 3 , but a relatively high collision frequency, v being around 10 6 /s. In con-
sequence, the refractive index is insufficiently low to cause reflection while the
attenuation at the lower end of the HF band is quite high.
The Kennelly-Heaviside or E layer occurs at around 110 km and is densest
during the day. It has an electron density TV of around 10 u /m 3 and a collision
frequency v below 10 5 /s; thus it is a reflecting rather than an attenuating layer. Its
critical frequency / c at midday is typically 3 MHz, but it decays to about 0.5 MHz
late at night. At large incidence angles the layer may reflect frequencies as high as
15-20 MHz during the day. Note that, because of the curvature of the earth, a wave
leaving the earth tangentially cannot reach the E layer at more than about 80°
incidence.
The Apple ton or F layer is the highest of the three, and since it is the most
exposed to solar radiation it is the most densely ionised. The ionisation persists
666 High Frequency Antennas

through the night at an altitude of around 250 km, During daylight hours the layer
intensifies, thickens and usually develops two distinct maxima, the ¥x layer at
around 200 km and the F 2 at around 320 km altitude. The electron density is
around 3 x 1011 /m 3 , but the collision frequency v is very low, less than 10 3 /s. The
critical frequency fc is thus around 5 MHz while the attenuation is negligible. The
F layer is the most important for long-range communication as it has the highest
critical frequency and provides the longest 'hops', The F layer remains usable dur-
ing the night although its critical frequency is lowered as the ionisation decays.
Fig. 16.3 shows in diagramatic form the reflection of an HF wave by an ionised
layer. For any given critical frequency and radio frequency there is a minimum
incidence angle given by eqn. 16.3, and this determines the minimum single-hop
sky wave range or 'skip distance'. For any given range, there is a maximum radio
frequency given by the same equation; this is called the 'maximum usable frequency'
or MUF.

skip distance

Fig. 16.3 Skip distance

Fig. 16.4 shows a typical relationship between the single hop range and the angle
of elevation at which the waves leave and return to the earth. It is advantageous to
design the antenna to fire at the optimum elevation angle for the desired path, but
in practice there is no single precise optimum angle because of the varying heights
of the ionised layers. The design angle of fire is thus a compromise, but the eleva-
tion beamwidth is usually wide enough to cover the variations.
Over long paths the wave may be reflected several times from the ionosphere,
bouncing off the sea or land between each hop. If the frequency is well below the
MUF for the route, several different paths may be possible, having various numbers
of hops and using the E, ¥x or F 2 layers or perhaps a combination of reflections
from two different layers (Fig. 16.5). It is even possible for a superior reflection
from an E layer to occur between two F layer reflections; this is known as an M
reflection. Since multipath propagation introduces distortion due both to wave
interference and to differential path delay, it is usual to minimise these effects by
working on a frequency fairly close to the MUF so that only one mode of propa-
gation can exist. The use of the highest possible frequency also minimises attenu-
ation. To allow for unpredictable variations in the MUF itself the optimum working
frequency is usually taken to be 85% of the MUF forecast.
Fig. 16.6 shows a typical MUF prediction for a particular route. The curves take
into account a number of different effects:
High Frequency Antennas 667

30-
F2 320 km
F 250km
F1 200 km
.E 110 km

range, km 2000 4000

Fig. 16.4 Elevation angles for skywaves

Fig. 16.5 Multimode propagation

SUMMER WINTER
50

0 hrs 12 24 0 hrs

Fig. 16.6 Maximum usable frequencies


668 High Frequency Antennas

(i) A day/night variation, MUF being higher during the day


(ii) A summer/winter variation, MUF being higher during the winter
(iii) A geographical variation since solar radiation intensity varies with latitude
(iv) A sunspot variation, MUF being higher during sunspot maximum.
The other major constraint on choice of frequency is D-layer absorption, which
determines the lowest usable frequency or LUF. Because of the variations in both
MUF and LUF, several different operating frequencies are needed to maintain
communication on any given route for the maximum possible proportion of time.
It is not possible to achieve 100% availability, as during magnetic storms following
solar flares the LUF may actually exceed the MUF, but with proper equipment and
correct choice of frequency 98% availability can be attained.
Although the ionospheric sky wave is the dominant HF propagation mode, it
should be noted that ground-wave propagation is also possible. The ground wave
which most commonly occurs at HF is not a true surface wave and is not bound to
the surface in the same way as a microwave surface wave over a corrugated or
dielectric-clad conducting sheet. This ground wave is vertically polarised, and is
automatically launched by any vertically polarised radiator having a significant
field component at zero elevation angle. Over sea, which has a good conductivity,
the wave propagates with the expected inverse-square-law power decay out to
around 50 km range, after which the curvature of the earth becomes significant and
the rate of decay increases. Since the velocity of the wave in the sea surface is
slightly less than that in free space, the phase fronts are tilted slightly forwards as
in Fig. 16.7 and the direction of propagation is curved downwards. This gives some
beyond-the-horizon radiation, although with a higher rate of decay than the free
space wave.

wavefronts radio horizon

Fig. 16.7 Ground-wave propagation at HF

Over land which has poor or very poor conductivity, the ground wave is greatly
attenuated and dies away within a few tens of kilometres of the launching antenna.
There is then a silent zone which extends out to the first sky wave skip distance.
This section has shown that the propagation of HF is fundamentally different
from that of any other frequency band. Ionospheric physics is a major subject in
its own right, and the outline given is necessarily brief, being intended only as an
introduction to antenna design. A far more detailed approach may be required
if any unusual HF system is contemplated, and References 28 and 29 may prove
helpful.
High Frequency Antennas 669

16.1.2 Characteristics ofHF signals


As a result of the use of the ionospheric sky wave, high-frequency signals have a
number of special characteristics which in turn affect the antenna requirements.
First, the waves are distorted because the ionosphere is neither regular nor
smooth. The ionised layers vary in altitude and density and may sometimes be
appreciably tilted. In consequence, the reflected waves are no longer plane, but
distorted so that they appear to come from a zone some 2° wide rather than a
point source. In elevation the spread may be considerably greater, particularly if
several different modes are present as in Fig. 16.5. This distortion of the wave-
fronts limits the receiving gain which can be used effectively, but as practical site
and cost constraints usually predominate, this limitation is rather academic.
Since signals following different paths arrive at different times, multipath dis-
tortion may occur if more than one mode is present. The polarisation of the reflec-
ted waves may also be distorted. It can be shown that the refractive index of an
ionised layer in the presence of the earth's magnetic field is slightly different for
right and left hand circularly polarised waves. Any wave incident on the layer can
be resolved into two opposite circularly polarised components, which then follow
slightly different paths. If both are reflected, they then recombine with a random
differential phase error which causes the resultant polarisation to vary unpredictably.
The combination of ionospheric roughness, multipath propagation and polari-
sation distortion causes variations in the received signal level. These variations
have a periodicity ranging from a fraction of a second to a few seconds, and are
known as 'fast fading'. The more gradual variations produced by the formation and
decay of the ionised layers are called 'slow fading'.
Fast fading may be countered by diversity reception. Two or more substantially
independent samples of the incoming wave-fronts are received, and combined in a
suitable processor. Because the fades in the separate channels are substantially
uncorrelated, the percentage of time for which at least one channel is providing a
satisfactory signal rapidly approaches 100%. For example, with two completely
uncorrelated 90% channels, a 99% probability of a satisfactory signal level is obtained.
Several different kinds of diversity are possible, using different methods of
obtaining substantially uncorrelated signal channels:
(i) Space diversity in which two receiving antennas are spaced ten or more wave-
lengths apart
(ii) Polarisation diversity in which two orthogonally polarised receiving antennas
are used
(iii) Angle diversity in which two receiving beams select different arrival angles so
as to separate the modes
(iv) Frequency diversity which uses simultaneous transmission of the same message
on two separate frequencies
(v) Time diversity, in which the message is transmitted twice with a short delay
between to decorrelate the fading. Only one receiver is needed, but the infor-
mation rate is halved.
To supplement diversity, automatic error correcting equipment may be used.
670 High Frequency Antennas

Signals are coded and a parity check is used to identify distorted groups so that a
repeat may be requested automatically. Telegraphic codes have a second advantage
in that they make optimum use of the limited HF spectrum which is available at
any given time.
These techniques, used with precise frequency control, enable many simultaneous
telegraph signals to be packed into the spectral band which would be needed for a single
speech channel. Note that multipath distortion limits the speed of telegraphy at HF.
It will be seen that, although wideband antennas are needed to accommodate the
slow changes in the ionosphere, the instantaneous bandwidth of any individual
signal channel may be quite small. However, one antenna may be used to transmit
or receive on several frequencies simultaneously.
High-frequency systems differ from microwave systems in that they are almost
always limited by external rather than internal noise. The thermal noise in a perfect
receiver is — 204 dBW in a bandwidth of 1 Hz. In a hypothetical receiver of overall
noise figure 6 dB and bandwidth 6 kHz the thermal noise is then — 160 dBW. At
HF, the background noise picked up by the antenna is normally much higher than
the thermal noise, and might typically be —130 to —100 dBW in the 6 kHz receiver.
The external noise is made up of contributions from several different sources:
(i) Manmade noise from vehicle ignition systems, industrial machinery, power
lines and similar sources
(ii) Manmade interference from other radio services, including spurious trans-
missions outside nominal allocated spectra
(iii) Natural noise from thunderstorms
(iv) Extraterrestrial noise from the sun or from the galaxy.
Since noise as well as signals may propagate by sky waves, noise sources such as
tropical storms may make an important contribution to the total, although they
may be thousands of kilometres distant.
Because the HF spectrum is external-noise dominated, receiving antenna direc-
tivity is more important than absolute gain. A lossy but highly directive receiving
antenna may give a better signal/noise ratio than a lossless antenna of lower gain.
For transmitting, however, maximum power gain is required, and directivity is only
important from the point of view of reducing interference with other users.
Because of the changing quality of the ionospheric path, HF skywave losses may
vary greatly. Under favourable conditions the ionosphere is a near-perfect reflector,
and at the higher end of the band where the D layer loss is small, communications
may be possible over antipodal paths with only a few watts of transmitted power.
The downward wave from the ionosphere may sometimes be focused into a small
area by the curvature of the reflecting layers, and then an artificially high signal
may be received (Fig. 16.8). On the other hand, under unfavourable conditions, it
may be necessary to employ very high powers and high antenna gains in order to
maintain communication.
While the familiar concepts of antenna gain and inverse-square-law power decay
apply equally to HF as to other bands, it should be noted that the method of calcu-
lating system parameters in practical applications may be somewhat different.
High Frequency Antennas 671
For system design purposes, a standard transmitter is often assumed, having a
power of 1 kW and driving a short vertical monopole antenna of gain 4.7 dB over
isotropic. The signal strength S in volts/metre at a range r metres from the antenna
is then given approximately by
300
S = (16.6)

assuming a perfectly conducting flat earth. At 1 km distance the field strength is


300 mV/m. Some propagation forecasts calculate the field at the distant receiver
assuming the above standard transmitter, taking into account D layer as well as
inverse-square-law losses.

irregular
ionosphere

Fig. 16.8 Ionospheric focusing


SUNSPOT MINIMUM SUNSPOT MAXIMUM

40-
20 MHz

20
15 MHz

T3

hrs 12 24

Fig. 16.9 Signal/noise ratio

These forecasts also estimate the noise level at the distant site assuming an omni-
directional receiving antenna and a receiver of standard bandwidth, perhaps 6 kHz
or 1 kHz. The result is presented to the system designer in the form of a chart of
expected signal/noise ratio against time of day for various available frequencies. It
is then relatively easy to determine how much additional transmitter power, or
transmitting or receiving antenna gain, is needed to provide the quality of service and
fading resistance specified. An example of a chart of this kind is shown in Fig. 16.9.
672 High Frequency Antennas

Note that the communication engineer still occasionally uses a horizontal dipole
at optimum height as an alternative gain standard. Its gain is 8.15 dB with respect
to an isotropic antenna.
Because of the relative unimportance of small losses in an external-noise-limited
receiving system, the impedance match of receiving antennas is not critical, and
VSWRs up to 3.0 are usually acceptable. For transmission a good match is needed
to protect the amplifier, and while a transmitter can usually tune out mismatches
of up to 2.0, VSWRs of 1.5 or less are generally preferred, particularly when the
antenna and transmitter are separated by several wavelengths of feeder line.

16.2 The ground at HF

16.2.1 Ground reflections


The conductivity of perfectly dry soil, sand, rock or concrete is very poor. Sand,
for example, may contain a substantial proportion of quartz, which is an excellent
insulator. In nature, however, the ground is seldom perfectly dry, and even when
the surface is dry, moisture is often present underneath. Pure water is also a very
poor conductor, with a conductivity of around 0.0001 S/m, but water with a trace
of salt contains free ions and its conductivity is much higher. The conductivity of
normal soil thus arises principally from its impure water content, and varies in the
range 0.001 to 0.05 S/m according to the amount of water present and its degree
of salinity. The conductivity of a beach may be in the range 0.1 to 1 S/m. The
conductivity of sea water is around 5 S/m, while a strong salt solution may have a
conductivity of 15 S/m or more.
The dielectric constant of the ground is also affected by the moisture content.
Dry soil, sand and rock have dielectric constants of 4 to 7, while water has a dielec-
tric constant of 80. The dielectric constant of normal moist soil varies between 10
and 20 according to the amount of moisture present.
For the engineer who likes to calculate everything precisely, natural ground at
HF is a most unsatisfactory material. Not only does its composition and moisture
content vary with location and depth, but the moisture content also varies from
day to day according to the weather and the natural height of the water table
within the soil.
The penetration of HF waves into the ground varies greatly with conductivity.
In really dry ground the waves penetrate down to 100 m or more, while in wet
ground the penetration depth is only a few metres. In pure fresh water the penetra-
tion depth is around 10 m, but in sea water it is only a few centimetres. Wet layers
at a very considerable depth below dry soil may thus play a significant part in
determining the overall behaviour of the ground.
The behaviour of uniform ground at HF can be deduced from a constant e called
the complex dielectric constant and defined as
e = k-60jo\ (16.7)
High Frequency Antennas 673
where k is the dielectric constant (equal to unity in free space), o is the conduc-
tivity in S/m and X the wavelength in metres.
Putting in the numerical values quoted earlier, it will be seen that in sea water
the conductivity term predominates, and it is therefore possible to regard the sea as
a rather lossy conducting sheet. All the other surfaces have complex dielectric con-
stants in which both the real and the imaginary parts are significant, and in con-
sequence the behaviour of the ground at HF cannot be represented either by a
simple dielectric or by a simple conducting-sheet model.
For horizontal polarisation, or more correctly for polarisation perpendicular to
the plane of incidence, the complex voltage reflection coefficient Rh of the ground
is given by
cosfl - ( e - s i n 2 f l ) 1/2
(16.8)
cos0 + ( e - s i
where 6 is the angle of incidence and e the complex dielectric constant as defined
above.
For perfectly conducting ground Rh is equal to — 1 for all angles of incidence.
For waves incident at low angles on normal ground, R^ is approximately equal
to-1.
A horizontal current element above the ground thus has a reflected image in the
ground in which the current flow is in the opposite sense and the ground-reflection
reinforcement occurs at angles of fire for which the path difference is a half wave-
length. Fig. 16.10 shows the reinforcement of a horizontal radiator at height h by
its image in a perfectly conducting ground. The angle of fire a is given by
Ah sin a = X (16.9)

high beam

Fig. 16.10 Direct and ground-reflected waves

If the ground is not a perfect conductor, the reflection coefficient at large


incidence angles (small values of a) has a smaller value but the phase change remains
substantially 180° so that the angle of fire a is scarcely changed (Fig. 16.11). The
reduced reflection coefficient may lead to a small loss of gain, usually less than ldB.
674 High Frequency Antennas

At small incidence angles (large values of a) the reflection coefficient falls and the
loss may be correspondingly greater, up to 3dB; but even here the gain remains
some 3dB higher than it would be if the ground reinforcement were absent. For
practical purposes, therefore, and particularly for antennas having low angles of
fire, the effects of ground imperfections may often be ignored if the polarisation is
horizontal. The majority of large HF antennas use horizontal polarisation for this
reason.

horizontal polarisation

0-5-

angle of 30° 60° 90°


incidence
Fig. 16.11 Reflection coefficient of the ground
Dielectric constant = 10
Ground conductivity = 10 m S/m
Frequency = 18 MHz
e= 10-/10

For vertical polarisation, or polarisation in the plane of incidence, the complex


voltage reflection coefficient Rv of the ground is given by
e cos 6 —(e — sin20)1/2
(16.10)
ecosfl + ( e - s i n 2 0 ) 1/2
For highly conducting ground, e is large and Rv tends to unity. A vertical cur-
rent element above the ground has a reflected image in which the current flow is in
the same sense, so that reinforcement occurs at low angles of fire (Fig. 16.12). This
situation can only occur in practice over ground which has been extensively improved
with screening mesh, or over sea. Ships and coastal radio stations commonly use
vertically polarised antennas to obtain low angles of fire.
When the conductivity is finite, eqn. 16.10 reveals a 180° phase change at
large incidence angles as the numerator tends to zero. Over sea, this happens at
incidence angles greater than 89° and the formation of the low-angle beam is
scarcely affected. Over land, the phase change occurs at incidence angles around
75°, and the reflection coefficient falls to a low value in the region of the change
High Frequency Antennas 675
(Fig. 16.11). The angle at which this occurs is called the Brewster angle. Antenna
performance is seriously affected, particularly as the minimum reflection occurs at
an angle which is comparable with the most probable arrival angle for HF skywaves.

high beam

low beam

Fig. 16.12 Radiation from a vertical antenna

At low angles of fire, the ground reflection tends to cancel the direct wave, just as
it does for horizontal polarisation. Thus the efficiency of a vertically polarised
antenna over natural ground is poor at low elevations and varies unpredictably with
ground moisture content. The use of vertically polarised HF antennas over unim-
proved ground is generally avoided when a low angle of fire is required.

16.2.2 Ground improvement


It has been shown that the radiation patterns of vertically polarised antennas are
degraded over unimproved ground. Ground improvement is expensive, and it is
necessary to find a cost-effective compromise between ground improvement and
the alternative of adding more transmitter power or antenna aperture to restore
overall performance. Where possible, of course, the problem is side-stepped by
using horizontal polarisation.
The overall size of the improved area depends upon the angle of fire required.
Theoretically the improvement should cover the first Fresnel zone, but since this
zone may extend for several kilometres in the direction of fire the cost of this may
be prohibitive. It has been shown by Bach Andersen1 that a much smaller improved
area is sufficient to restore most of the lost performance. In one example improve-
ment to one quarter of the first Fresnel zone resulted in a loss of gain of only 3dB,
as compared with 14dB for unimproved ground. This suggests that a suitable
practical length L of the improved zone in the direction of fire might be

L = (16.11)
8 ( 1 - c o s a)
where a is the angle of lire as before. In other directions improvement to a distance
of \/4 from the antenna is suggested (Fig. 16.13). The lateral extent of the improved
zone depends very much on the type of azimuth coverage required (see Wait26).
Ground improvement may be by parallel or radial wires when the ground-current
676 High Frequency Antennas

flow is known to be unidirectional, or it may be by rectangular wire mesh when the


direction of current flow is unknown or is frequency dependent, as for example
underneath a logarithmic antenna. Rectangular wire mesh is obviously more expen-
sive, since it not only requires twice as much metal, but also introduces jointing
problems.

Fig. 16.13 Ground plane for vertically polarised antennas

From microwave experience, it might be expected that a mesh size of around


0.05X would form an adequate reflecting screen. At HF, however, the diameter
of the wire is so small that the inductance of the mesh becomes appreciable and it
is not advisable to make the mesh size much greater than 0.0IX. Note that the
behaviour of a mesh is not the same as that of two orthogonal but mutually insu-
lated sets of parallel wires, since in the mesh circulatory currents are possible. Note
also that the presence of the ground with its high dielectric constant affects the
behaviour of the mesh if the mesh is actually laid upon or buried beneath the sur-
face. A more efficient although less convenient arrangement is to support the mesh
at a height of A/4 above the ground so that the evanescent field beneath the mesh
cannot excite the ground unnecessarily.
Recent years have seen enormous increases in the computing power available
to engineers, and there is now justification for much more sophistication in the
design of ground screens. It should now be possible to optimise the size and shape
of the screen and the spacing and orientation of the wires so as to obtain not only a
cost-effective increase in forward gain, but also a significant improvement in side-
lobe and backlobe performance. Note that in wideband antennas the size of ground
screen required is frequency dependent; thus the wire spacing or mesh size can be
increased considerably in the outer zones. With a logarithmic antenna, a logarithmic-
ally periodic ground screen may be appropriate. However, there is a practical limit
to the refinement which can be incorporated into the screen, and coarse quantisa-
tion into zones of different mesh size or wire spacing is convenient for construction
and installation.
High Frequency Antennas 677
The choice of material for ground screens requires careful consideration. Most
metals have only a limited life because of corrosion, and copper remains one of the
best materials in spite of its high cost. One of the likely applications of ground
screens is in linking coastal antennas to the permanently wet part of the beach.
Here aluminium and steel are particularly vulnerable to corrosion and can have only
a short life if unprotected. Stainless steels are not normally resistant to sea water,
and in addition have a relatively low conductivity. Unorthodox methods of ground
improvement such as keeping the area permanently waterlogged should not be
forgotten in these circumstances.

16.3 Non-resonant HF antennas

16.3.1 Long wire antenna


This section is concerned with travelling-wave antennas. In a travelling-wave antenna
the wave propagates along the radiator in one direction only, any remaining power
at the far end being absorbed in a matched load so that there is no reflected wave.
A long terminated wire radiator is non-resonant and its input impedance varies
only slightly with frequency.
Consider the radiation pattern of a long wire length L. The phase varies along
the wire at the same rate as that of a free-space wave, and at small angles 6 from
the axis of the wire in the direction of propagation contributions from all the
elementary portions of the wire are still nearly in phase (Fig. 16.14). Along the

Fig. 16.14 Radiation from a long wire

direction of the axis the field must be zero, since the radiation pattern of each
elementary portion includes a sin 6 term. When the angle becomes substantial, the
contributions received from all parts of the wire differ in phase by:

If the amplitude of the wave is assumed to be uniform along the wire, that is, if
the loss of energy by radiation is neglected, then a null occurs in the pattern when
this phase difference is equal to any integral multiple of 2n. A maximum occurs
approximately halfway between each null. The radiation pattern of a uniformly
energised wire antenna of length 3X is shown in Fig. 16.15. Note that the largest
lobe is nearest to the direction of wave propagation along the wire.
678 High Frequency Antennas

It can be shown that the radiation pattern of any uniformly energised termin-
ated long wire is given by
sin 9 \nL (1 — cos0)
(16.13)
1 - cos 0 sin 'L X

The position of the first maximum is given approximately by

cos<9 = 1 - 0 . 3 7 — (16.14)

and this function is plotted in Fig. 16.16 for typical practical values ofX/L.

Fig. 16.15 Radiation pattern of an end-fed wire

40

20-

Fig. 16.16 Radiation angle for a long wire

The assumption that the excitation is uniform is not justifiable in practice, since
perhaps half the power may be radiated, but the effect of this approximation on
the position and shape of the pattern lobes is relatively small. Since the main lobe
of the pattern in three dimensions lies on a cone surrounding the axis of the wire,
the long wire antenna is of little use by itself. It becomes useful when the radiation
patterns of two or more identical long wires are combined to form a single beam,
as described in the following sections.
High Frequency Antennas 679
16.3.2 Vee antennas
The inverted-vee antenna described by Bruce5 is a simple practical configuration
based on the long wire. It is inexpensive since it requires only one mast (which must
be non-conducting) and one wire radiator terminated in a resistive load at the far
end. Fig. 16.17 shows a typical design. The angle of inclination, 30°, is chosen to
make the maximum radiation from each section approximately horizontal (Fig.
16.16). This also ensures that the phase reversal brought about by the fact that one
half of the wire goes up and the other half down is compensated by the phase
reversal produced by the distance between the centres of the wires, approximately
0.5X less than the length of the current path.

3OO/i

2-5*

Fig. 16.17 In verted-vee antenna

The Bruce antenna is vertically polarised and therefore requires an improved


ground surface for efficient operation. A good ground connection is needed for
the resistive termination, which is likely to be around 300 £1 but is best adjusted
experimentally to take account of ground imperfections and other secondary effects.
If used for transmission, more than one wire conductor may be used to reduce
voltage stresses and improve the impedance match.
The bandwidth of the Bruce antenna is about an octave. Its radiation patterns
show large sidelobes near to the main beam and at high elevations, as might be
expected from consideration of the individual patterns of the up and down con-
ductors. The efficiency of the Bruce is low, as around half of the total input power
may be wasted in the matched load.
The Bruce has an unbalanced input, and generally requires a matching trans-
former to couple it either to a coaxial or a balanced twin transmission line. Its
direction of fire can be reversed by interchanging the feeder and the load.
A second type of vee antenna is described by Watt-Carter and Young.27 The
sloping-vee is shown in Fig. 16.18. Like the Bruce antenna, it requires only one mast,
but unlike the Bruce, it is a balanced antenna and it is horizontally polarised. The
angle between the descending wires is chosen so that the principal lobes of the two
radiators are superimposed. Referring to Fig. 16.16, an included angle of 45° is
correct for wires of length 5X. The mast height is chosen to make the antenna tilt
angle the same as the design angle of fire of 15°. Since most of the radiated power
is reflected from the ground, the ground should be reasonably level and free from
obstructions, but ground conductivity is not as important as it is for the Bruce
antenna because of the use of horizontal instead of vertical polarisation.
The radiators are terminated in matched loads of around 800 £1. These loads
680 High Frequency Antennas

must be earthed, and the antenna is generally used only for reception because of
the difficulty of making satisfactory high-power earthed terminations. Like the
Bruce, the sloping-vee antenna has an octave bandwidth, substantial sidelobes, and
a low efficiency. The antenna has a high input impedance, around 800 £2, and is
conveniently fed by a tapered line transformer from a balanced twin transmission
line.

direction

load

F ig. 16.18 Sloping-vee an tenna

16.3.3 Rhombic antenna


The rhombic antenna, also due to Bruce,5 combines the principles of both of the
vee antennas previously described. It consists of four radiating wires of equal length
suspended between four masts to form a horizontal rhombus. A typical antenna is
shown in Fig. 16.19.

Fig. 16.19 Rhombic antenna

The antenna is balanced and radiates horizontal polarisation, and like other
travelling wave antennas it requires a matched load at the far end. If the load is
not correctly matched a backlobe is formed by the reflected wave and the input
impedance is degraded.
Referring to Fig. 16.20, each of the four radiators generates a conical principal
lobe at an angle of about 29° to its axis. If the rhombus vertex angle were 58°,
these four lobes would all combine along the longitudinal axis of the antenna, the
half-wave fore-shortening of the path between the forward and rearward portions
High Frequency Antennas 681
compensating for the phase reversal due to the opposite inclination angles. An
improvement in gain can be obtained by choosing a slightly smaller vertex angle,
say 55° in the above example. This causes the single beam to break up into two
lobes corresponding to the intersection of slightly overlapping cones. One beam is
at around 15° elevation and is precisely in the wanted direction, while the second
beam at — 15° reinforces the first after ground reflection, the height of the rhombic
above ground being calculated from eqn. 16.9, so that the direct and reflected
beams are in phase.

Fig. 16.20 Radiation from a rhombic antenna

The above is probably a sufficient explanation for most practical purposes. A


rigorous design theory would take into account the effect of the amplitude decay
along the wires as power is radiated, the mutual impedance between the radiators
and the effects of ground imperfections. Since a great deal of practical design data is
already available in the literature and since a large number of proven standard
designs are in current operation, there is no need to include a comprehensive analy-
sis here. A useful design chart is given by Norman and Ward.19 Fig. 16.21 gives the
optimum vertex angles for rhombics of various sizes, together with the gain which
can be achieved if the height above ground is such that the direct and reflected
waves are in phase.
Note that a rhombic optimised for a single frequency and a given angle of fire
retains its optimum performance over a comparatively narrow band. However, it
continues to provide a useful performance over a bandwidth of around an octave,
its match in particular remaining reasonably good. Less satisfactory is the sidelobe
performance, since a substantial proportion of the total input is wasted in sidelobe
radiation, particularly when working at frequencies remote from the design optimum.
Because of this sidelobe radiation and also because of the power loss in the matched
termination, the power gain of the rhombic is much lower than its directivity gain.
The latter can be as high as 25dB over isotropic, giving the rhombic a limited azi-
muth sector coverage which restricts its use to point-to-point and other narrow-
sector applications.
The rhombic is relatively insensitive to ground imperfections, thanks to its
horizontal polarisation, and it does not require any earth connections.
The impedance of the rhombic is around 600-800 £2 depending upon its size
and design. It is conveniently matched into a twin transmission line by a short
682 High Frequency Antennas

tapered-line matching transformer. The load impedance is best adjusted on site.


The direction of radiation can be reversed by interchanging the feeder and load.
The rhombic can be used to handle high powers, and it is then usual to employ
multiple-wire or cage radiators instead of single wires. Multiple-wire rhombics also
have a better impedance match over the band.

10
side length, I
Fig. 16.21 Vertex angles and gains for rhombic antennas of optimum design

The large rhombics used for long-range communications have small vertex
angles and their resultant pattern lobes, corresponding to the direct and ground
reflection beams, cannot lie far from the horizontal plane. It is not possible to
design large rhombics to produce an optimum performance on high-angle short-hop
paths. The largest rhombics, of around 10A side, are at their best at elevation angles
of fire around 5°, for which masts of 3X height are necessary. A 5X rhombic is
suitable for 10°, a 3A rhombic for 15°. For each of these elevation angles, there is
little increase in gain if the rhombic is made larger than the suggested size.

16.3A Beverage antenna


Although superficially a travelling wave antenna, the antenna described by Beverage,
Rice and Kellogg3 works on an entirely different principle from the others in this
section. The Beverage antenna consists essentially of a very long wire supported on
insulating poles a small fraction of a wavelength above the ground. Its length is of
order 10 wavelengths and it is mainly useful for reception as its efficiency is very
low. Its usual application is at LF where the choice of antennas is more limited
but it may occasionally be used at the lower end of the HF band.
To understand the Beverage antenna it is necessary to consider the distortion of
a wave front travelling over the earth's surface. The wave front is not perpendicular
High Frequency Antennas 683
to the ground, but is tilted as a result of the slowing down of the wave at the sur-
face. Thus, a vertically polarised wavefront generates a horizontally polarised
longitudinal component which excites a small voltage in the long-wire antenna.
Since the free-space velocity of the wave and the velocity of the induced wave in
the wire are almost the same, the wave in the antenna gradually builds up along its
length until it reaches a usable signal level. The antenna is directional, since if the
wave is propagating at an angle to the wire the distance over which reinforcement
can take place is limited to a fraction of the total aperture. A matched load at the
remote end of the antenna absorbs any signals picked up in the unwanted direction.

16.4 Resonant HF antennas

16.4.1 Monopole an tenna


The basic theory of monopoles has been covered in Chapter 15. At HF the useful-
ness of monopoles is limited by the wide operating bandwidths usually required.
A simple HF monopole consisting of a rigid quarter-wave conducting tube
mounted over a perfect ground has an impedance of around 30 12 and radiates
vertical polarisation with a maximum at zero elevation. For spot-frequency opera-
tion on a coastal site such an antenna would be satisfactory, but on poorly conduct-
ing sites its low-angle coverage would be limited by the amount of ground improve-
ment undertaken.
At frequencies below resonance, the radiation resistance of a monopole decreases
and series inductance or top capacitance has to be added to maintain a non-reactive
impedance. The radiation pattern is little changed, but the combined effects of
reduced radiation resistance and resistive losses in the matching reactances result in
a decreasing overall efficiency as the frequency is lowered.
At frequencies above quarter-wave resonance, the radiation resistance increases
and the opposite reactance compensation is needed to maintain a purely resistive
impedance. As the resonance gradually changes from quarter to half-wave, the
elevation pattern coverage is reduced but the maximum remains as near to zero
elevation as the extent of the ground improvement permits. Above half-wave
resonance, at which the input impedance becomes very high, the elevation pattern
breaks up into two or more lobes.
Since bandwidths of up to a decade may be required at HF, the large impedance
variations and the elevation pattern break-up of the monopole mean that several
would be needed to cover the whole band. These would need to be spaced suffici-
ently far apart to avoid mutual coupling and consequent pattern distortion, and so
would occupy as much space as a single wideband antenna. Thus, the simple mono-
pole is not widely used at HF, but it forms a starting point for other designs.
When elevation pattern break-up rather than impedance match is the limiting
factor, as is the case in most receiving applications, the elevated-feed monopole
provides a useful improvement (Fig. 16.22). By feeding the monopole about
684 High Frequency Antennas

one-third of the way up rather than at ground level the break-up of the antenna
currents into higher-order modes can be delayed and the elevation coverage can
be preserved.

insufator
transformer

feeder

Fig. 16.22 Elevated-feed monopole

The antenna consists of two concentric tubes, the larger forming the lower part.
At the feed point, a small transformer may be used to give the best compromise
match into the coaxial-cable feeder, which passes up through the lower tube. In
practice, elevated-feed monopoles may be used over bandwidths of up to 4:1 so
that two monopoles are sufficient to cover the entire HF band.
When impedance bandwidth is the limiting factor, an improvement can be
obtained by increasing the monopole diameter. If a wire-cage rather than a solid-
tube construction is employed, the monopole diameter may be made very large and
a working bandwidth exceeding 2:1 is readily obtained. In one practical design
described by Mason16 a double cone monopole is used to obtain a bandwidth of
2.6:1 with a VSWR not exceeding 1.6 into a 50 Q feeder. Fig. 16.23 shows a typi-
cal practical double-cone monopole antenna. When the limiting factor is physical
size, as it may be at the lower end of the band where a quarter wavelength may be
25 m or more, other variations may be used. Fig. 16.24 shows the inverted-L and T
antennas, in which a horizontal wire provides top capacitance and enables the
height to be reduced by an amount roughly equal to the length of the horizontal
portion. Since most of the current flow is in the vertical portion, the radiation
pattern remains substantially vertically polarised. Although the height is reduced,
two supports are now required so that there is little cost saving unless suitable
supporting structures are already available.
Top capacitance may also be added in the form of an umbrella consisting of the
upper portions of a circle of guy wires, Fig. 16.25.
Series inductance loading is convenient in practice because it is easily incorpor-
ated in self-supporting radiators. The position of the inductance or inductances
affects both the impedance and radiation pattern behaviour particularly at the
upper end of the working band where break-up is approaching.
High Frequency Antennas 685

\
unbalanced feed

Fig. 16.23 Double-cone monopole

feed feed

inverted-L T-antenna

Fig. 16.24 Inverted-L and Tantennas

top cap
mast

Fig. 16.25 Umbrella antenna


686 High Frequency Antennas

For mobile applications, loaded flexible 'whips' may be used. These are dis-
cussed in Section 16.8.1.
Since external rather than internal noise predominates in HF communications,
the power gain of receiving antennas is often unimportant provided that the direc-
tivity and linearity characteristics are acceptable. Thus it is becoming increasingly
common to use short low-gain monopoles with integral low-noise amplifiers in
place of full-size monopoles. These 'active' monopoles occupy less space and may
be considerably cheaper to install. Note that the active monopole is not merely a
short monopole plus an amplifier, since the amplifier circuit may be designed so
that it also acts as an impedance transformer, matching into the very low impedance
of the short monopole and compensating for its frequency-dependent reactance.

16.4.2 Dipoles and slo t an tennas


Just as monopoles are usually employed as vertically polarised antennas, dipoles are
usually horizontally polarised. They are less sensitive to imperfections in ground
conductivity, but they must be raised well above the ground in order to have a
low angle of fire as discussed in Section 16.2.1.
Dipoles are usually operated at or near half-wave resonance, and their impedance
is consequently lower than that of open transmission line. A matching section may
be incorporated in the feeder, as in Fig. 16.26. In reception the bandwidth is

matching
section

Fig. 16.26 V'-matching to an antenna

limited by variations in the angle of fire, but for transmission the impedance is also
a critical parameter. The impedance bandwidth may be improved by reducing
the inductance/capacitance ratio in the same way as in the monopole, and designs
such as the 'bow-tie' (Fig. 16.27) and the cage dipole (Fig. 16.28) are often used
in practice in preference to single wires.
While it is possible to operate dipoles in full-wave as well as half-wave modes,
thus extending their coverage to a second frequency band, it should be noted that
their radiation patterns are then quite different. The use of harmonically related
frequency bands by amateur radio enthusiasts encourages this type of operation,
and it is even possible to operate on three bands with appropriate matching circuits.
Moxon18 describes one such design, but its application is limited to the special case
of harmonically related bands, and it does not meet the broader requirements of
commercial HF operation.
High Frequency Antennas 687
Slots are complementary to dipoles, and therefore in principle offer an alternative
family of antenna configurations. In practice slots are far less convenient at HF
because they need a large conducting sheet or mesh screen instead of a half-wave-
length conductor. Even when a conducting sheet is available, as for example the
side of a ship or the wing of an aircraft, it is not usually structurally permissible to
cut a slot large enough to resonate at HF. Nevertheless slot antennas remain an
interesting possibility for unusual applications.

feeder
Fig. 16.27 Bow-tie dipole

spreaders

feeder
Fig. 16.28 Cage dipo/e

Dipole and slot antennas are bidirectional unless a reflector is provided to


suppress the back radiation. A reflector may consist of a.second dipole, placed
around 0.2 wavelengths behind the first, but this arrangement is only satisfactory
over a relatively narrow band. A wire mesh reflector is better, since it remains
efficient over a wider band. However, the distance between the radiator and the
reflector, in terms of wavelengths, remains frequency dependent, so that the
reflector modifies both the forward radiation pattern and the impedance match as
the frequency varies.

16.4.3 Loop antennas


Loop antennas are useful for reception when space is very limited. In the loop
(Fig. 16.29) the voltages induced in the vertical sides are equal and opposite when
the plane of the loop is perpendicular to the direction of propagation of the inci-
dent wave, and differ by an amount proportional to the horizontal separation when
the plane of the loop is parallel to the direction of the incident wave. Since the
voltage is also proportional to the length of the sides, the total induced voltage
688 High Frequency Antennas

in a small loop (small in terms of wavelength, that is) is proportional to the area
multiplied by the number of turns.
The radiation pattern is thus the same as that of a short magnetic dipole, or a
short electric dipole with the electric and magnetic fields interchanged. It is given
by sin 0, where 6 is the angle between the direction of the incident field and the
axis of the loop.

incident
wave

Fig. 16.29 Loop antenna

Loops may be untuned, but they are far more efficient if tuned to resonance.
The inductance of a practical multiple-turn loop at HF is sufficient to make it self-
resonant, and a single-turn loop tuned by a small capacitance is more convenient
and more efficient. Its area may be considerably larger than that of a multiple-turn
loop if self-resonance is the limiting factor on loop area.

earthed
screen

balanced feeder

Fig. 16.30 Balanced screened loop

Apart from their use in domestic receivers, loops are mainly employed in direction-
finding (DF) applications. The design of DF antennas is a higly specialised subject
which cannot adequately be covered in this general chapter. However, it should be
noted that the reduction of errors in DF systems depends principally upon the
High Frequency Antennas 689
elimination of pick-up due to unwanted modes. The use of shielded or balanced
loops (Fig. 16.30) helps to reduce the sensitivity to unwanted (unbalanced vertical)
signals.

16.5 Directional HF arrays

16.5.1 End fire arrays


Endfire arrays provide a higher directivity gain than single resonant elements.
Unlike broadside linear arrays, endfire arrays provide increased directivity in both
planes, so that the elevation beamwidth is narrowed as well as the azimuth beam-
width. At HF, this means that a good low-angle elevation pattern can be obtained
without using excessively high masts. However, it is not easy to obtain the quality
of impedance match needed for high-power transmission, thus endfire arrays are
normally used only for reception.
An endfire array of horizontal dipoles is known as an HAD (horizontal array of
dipoles), Fig. 16.31. It consists often to twenty half-wave dipoles spaced rather less
than a quarter-wave apart and centre-fed from a common 600 12 balanced trans-
mission line. Its action is most easily explained in terms of transmission. The dipoles
are energised by a wave travelling along the line, each dipole extracting a proportion
of the power. The phasing is such that the radiated fields of the separate dipole
elements add naturally in the direction of fire. The dipoles load the line periodically
so that their reflections tend to cancel; thus the impedance bandwidth is greater
than that of an isolated dipole.

feeder load

Fig. 16.31 HAD antenna

At resonance, however, the dipole impedance is much lower than the impedance
of the transmission line. The first dipole thus extracts a large proportion of the
total power in the line. The second dipole extracts a large proportion of the power
remaining after the first, so that the resultant amplitude distribution along the
array is approximately exponential. The far end of the line is terminated in a
matched load, but little power remains in the line at the far end and the contribution
of the last few elements is minimal.
As the frequency is varied from resonance, the impedance of the dipoles rises
and the power extracted by the first few elements is reduced. Since more power
remains in the line, the contribution of the elements at the far end is increased, so
that the antenna directivity gain is increased. At the same time, the power wasted
in the end load is increased, so that the overall gain rises to maxima at the upper
690 High Frequency Antennas

and lower ends of the band where all dipoles are contributing and then falls away
steeply as the end load loss becomes dominant. The useful radiation pattern band-
width of the antenna is about an octave, but the impedance match is never particu-
larly good because of the overwhelming effect of the first dipole. In consequence,
the antenna is used for reception where impedance match is not critical.
An interesting variant of the HAD is the RCA Fishbone, (Fig. 16.32). This uses
ceramic insulators to connect the dipoles to the feeder, so that the capacitance of
the insulators controls the coupling. Less power is extracted by the first few dipoles,
enabling a more uniform distribution to be achieved with a consequent higher gain.
Since the dipoles all present a capacitive reactance to the line, the line has a reduced
phase velocity and the dipole fields no longer add in phase in the direction of fire.
The phase lag actually increases the gain slightly (see Hansen and Woodyard10). The
antenna is usable for reception over a band of up to two octaves.

IIUUUUUL
feeder
TTTTT/T T T T T TTT
coupling capacitors
load

Fig. 16.32 RCA Fishbone antenna

Both HAD and Fishbone antennas are slung horizontally between four masts or
towers. The elevation angle of fire is given by eqn. 16.9 in Section 16.2.1.
Vertically polarised endfire arrays are also possible but few have ever been used
in practice because of the need for extensive ground improvement if the potential
low angle of fire is to be realised.

T
4~
feeder
_L load

Fig. 16.33 Series phase array

One example was the series phase array described by Ladner.15 This was an end-
fire array of quarter-wave folded monopoles at quarter-wave spacing, Fig. 16.33.
The array was fed against the ground plane, and as in previous endfire arrays a
resistive load was used to terminate the far end, typically 300 O. A gain of around
12dB was claimed for an array four wavelengths long, but the impedance band-
width proved poor and the array was unable to handle high powers.
The series phase array required only two supports, and since these were parallel
High Frequency Antennas 691
to the radiators it was necessary to use wooden poles or towers rather than metal
masts.
All uniform endfire arrays can be made to fire on an opposite bearing by inter-
changing the feeder and the matched load, and this feature is useful in HF com-
munications when the same antenna is used at different times of the day for dif-
ferent circuits.

16.5.2 Broadside arrays


Broadside arrays provide better azimuth directivity for a given horizontal aperture
than endfire arrays, and they offer the additional possibility of beam steering by
phase variation.
Early broadside arrays were formed by folding or loading long wire antennas
in such a way that the radiated fields added in phase. These antennas were hung
from catenaries between substantial towers, and were generally called "curtain
arrays', However, correct phasing was only obtained over a relatively narrow band,
so that several separate curtains were needed to maintain service under all iono-
spheric conditions.
An obvious alternative is the broadside dipole or Koomans array, shown in its
original form in Fig. 16.34. It consists of rows and columns of A/2 dipoles. The
rows are spaced A/2 apart and fed by twisted vertical feeders; the currents in all
the rows are then co-phased. The dipoles themselves are end fed; this ensures that
all the dipoles carry equal currents even though their impedances (which depend
on their positions in the array) are different. To achieve unidirectional transmission,
a reflecting screen, or a second curtain of dipoles tuned to resonance, is spaced
A/4 from the driven curtain. The use of two identical curtains enables the direction
of transmission to be reversed. In normal use the two vertical feeders shown in
Fig. 16.34 are connected in parallel and carry co-phased currents. The direction
of the main lobe is then normal to the plane of the array. If the currents are deliber-
ately misphased the direction of the main lobe can be slewed by up to ± 15° from
the normal; this facility is useful when transmissions to several different directions
are required.
A special notation is used to describe curtain arrays of this type. In its most
general form the array is denoted by HRRS n/m/h where
H = horizontal polarisation
R = array with reflector
R = reversible array
S = array can be slewed
n = number of columns of A/2 dipoles
m = number of rows of dipoles
h = height of lowest row of dipoles above ground in wavelengths.
The array illustrated in Fig. 16.34 is of type HR 4/4/1.
Curtain arrays are extensively used for broadcasting. They have the advantage
that, by varying the numbers of columns and rows, their vertical and horizontal
radiation patterns can be varied independently according to operational requirements.
692 High Frequency Antennas

Thus the number of rows of dipoles depends on the distance of the service area
(Knight13) while the number of columns depends on the angle which the service
area subtends at the transmitter. The gain of a typical array is 22dB, relative to

t
2

to transmitter

Fig. 16.34 Broadside dipole array

that of an isotropic source in free space (22dBi). Fig. 16.35 shows gains of arrays of
type HR 4/m/h and Fig. 16.36 shows the elevation angle of the main lobe.
The type of curtain array described above can only be used for one broadcasting
band. Consequently large numbers of arrays had to be provided to cater for all
High Frequency Antennas 693
broadcasting bands and for all required directions of propagation. Modern curtain
arrays, however, can be used for up to four adjacent broadcasting bands, covering
a frequency range of nearly 2 : 1 . This is achieved by the use of centre-fed wide-
band dipoles, such as those shown in Fig. 16.28, and by branch feeding the rows of
dipoles, as shown in Fig. 16.37.

22 -

20

c
&
16

10
0.5 10 1.5 20
height of lowest row of dipoles above ground, A

Fig. 16.35 Gain of curtain arrays of type HR 4/m/h

The reflector curtain consists of a grid of horizontal wires spaced about 0.03A
apart. Althouth the array cannot be reversed, this restriction is not a serious limi-
tation because service areas seldom lie on reciprocal bearings. The number of
vertical feeders has been doubled, but this is an advantage because much wider
slew angles can be achieved by progressive phasing. Slew angles of up to 30° are
possible without a serious reduction of gain or the generation of strong sidelobes.
Horizontal and vertical radiation patterns of wide-band arrays, covering a 2:1
frequency range, are shown in Figs. 16.38 and 16.39.
Whereas at microwave frequencies it is usual to optimise phased arrays for some
relatively narrow bandwidth such as 10%, at HF the design is compromised to
permit operation over as wide a band as possible. When the frequency is raised
above the grating-lobe threshold, the gain reaches a plateau as the grating-lobe
loss compensates for the narrowing main beam. Provided that the element remains
workable, the array continues to provide gain in the wanted direction, but un-
wanted signals are now transmitted and received at wide angles rather than close
694 High Frequency Antennas

to the wanted direction of radiation. Thus, the array can be used successfully at
these higher frequencies provided the grating lobes can be tolerated, (see Fig.
16.40.)
50r

40

;30

o
£20
c
o
g
I
10

0 0.5 1.0 1.5 2.0


height of lowest row of dipoies above ground,A
Fig. 16.36 Elevation angle of main lobe of curtain arrays of the type HR n/m/h

At lower frequencies also, the degradation can be tolerated. The gain is deter-
mined by the overall aperture available regardless of whether the aperture is filled
or over-filled by elements. In receiving arrays the elements do not need to collect
all the power available, but merely to sample the field with sufficient element
gain to make external noise mask receiver noise; thus the mutual effects between
elements may be unimportant even with close element spacing.
Simple omnidirectional receiving elements such as elevated-feed monopoles
have the disadvantage of radiating equally on both sides of the array so that a
backlobe of equal magnitude to the wanted lobe is generated. Use of a reflecting
system enables the backlobe to be cancelled at the frequency for which the reflector
is optimised, but passive reflectors do not behave well over wide bands. Not only
does a reflector distort the forward primary pattern; it may even turn into a director
and reverse the pattern at some other frequency.
High Frequency Antennas 695
One solution is the use of doublet elements, fed in delayed antiphase so as to
give perfect cancellation in the backward direction. The elements are closely spaced
(Fig. 16.41), so that in the forward direction they do not add in phase except per-
haps at the highest frequency in the operating band. The resulting primary pattern
is a cardioid or quasi-cardioid, and considerable suppression of the backlobe is
obtained.

A
/2

Fig. 16.37 Wide-band curtain array


Dimensions shown are for mid-band frequency

A second solution, which has the additional advantage of permitting omni-


directional rather than sector scanning, is described by Morris er^z/.17 In the 'Medusa'
system 48 irregularly spaced omnidirectional elements were scattered over an area
696 High Frequency Antennas
so as to form a part-broadside part-endfire array. This array had both the forward
azimuth directivity of a broadside array and the elevation directivity and backlobe
protection of an endflre array. However, the shadowing of some elements by
0-67 f0 1 33 f0

(a) unslewed

(b) slewed

Fig. 16.38 Horizontal radiation patterns of wide-band HRS 4/m/h curtain array
f0 = design frequency
\30°
elevation
angle

20°

0.2 0.4 0.6 0.8


relative field strength
Fig. 16.39 Vertical radiation patterns of wide-band HR n/4/0.5 curtain array
f0 = design frequency

others meant that the maximum gain was less than that of a broadside array having
the same total number of elements; thus the average gain improvement was only
about lldB as compared with a single isolated element. The sidelobe and backlobe
High Frequency Antennas 697
performance was not particularly good, the sidelobes being more easily predicted
by statistics than by rigorous analysis.
Broadside arrays of more complex elements are possible, giving higher gains
but with limitations in scan coverage. Thus a broadside array of rhombics can
provide a very high gain, but is limited to a phase scan sector of a few degrees.

broadside beam

unwanted
grating
!obe

array elements

F ig. 16.40 Forma tion o f gra ting lobes

cancellation partial reinforcement

phase
reversal

Fig. 16.41 Directional receiving antenna

Inverted-vee elements can be used, giving a narrower coverage than monopolesbut


some backlobe protection, while logarithmic elements can be used to give very
good backlobe protection with forward coverage arcs of up to perhaps ± 50°.
Although individually wideband, logarithmic elements are degraded in both radia-
tion pattern and impedance match when too closely spaced, but they perform well
in the higher parts of the band when the array is in the grating-lobe condition.
Logarithmic antennas and arrays are discussed in full in Section 16.6.

16.5.3 Circular arrays


When all round coverage is required, for example for direction finding, a circular
array may be more cost effective than a number of linear arrays. A typical HF
circular array or Wullenweber may consist of 30-100 elements arranged in an equi-
spaced circle, fed from a central phasing device or goniometer (Fig. 16.42) The
elements may be monopoles, elevated-feed monopoles or active monopoles, and will
probably have associated reflectors to give a suitable broad directional primary
698 High Frequency Antennas

pattern. Alternatively, monopole doublets may be used to generate cardioid


patterns.
Usually only about one-third of the elements are used to form any given beam,
since the contributions of the back and side elements to the forward pattern are
insignificant. A relatively simple amplitude weighting may be used to improve the
side lobe performance.

fc feeders

goniometer

elements

Fig. 16.42 Circular array

The difficulty of the Wullenweber arises from the need to cover a wide band. A
single receiving array covering the whole HF band requires too many compromises,
but two arrays covering perhaps 2-8 and 8-32 MHz, respectively, are feasible.
The elements must therefore maintain an acceptable primary pattern over a 4:1
band, and must do this in the presence of the other elements. Possible problems
include the degradation of the primary pattern due to the reflector becoming a
director, or due to travelling waves running around the array. Fortunately these
effects can easily be checked by scale modelling before the full-size antenna is
constructed.

low band

high band

Fig. 16.43 Concentric circular arrays

If high- and low-band arrays are constructed concentrically to save space, and if
the high-band antenna is mounted outside the low-band circle to avoid obscuration,
there will be a wide difference between the gains at the upper and lower ends, and
High Frequency Antennas 699
the high-band array will have at least four times as many elements as the low-band
array (Fig. 16.43). If the high-band array is mounted in the centre of the low-band
array (Fig. 16.44), the number of elements can be the same for each. Obscuration
effects are limited to low arrival angles, and the same phasing system can now be
used for both bands.

! |
I I!| III '
high band

low band

Fig. 16.44 Concentric circular arrays

The circular array can only be correctly focused at one arrival angle. At higher or
lower elevations the pattern is defocused and there is a loss of gain accompanied by
a rise in near sidelobe levels. The effects are minimised by setting the phasing
system to give optimum focusing at medium arrival angles rather than at zero
elevation. The defocusing errors are then acceptably small for all but the highest
elevation signals.
The goniometer may be a mechanically rotating capacitive switch, or it may be
an all-electronic system. Future systems are likely to be of the latter type to over-
come the sampling-rate limitations of mechanical switches.

16.6 Logarithmically-periodic HF antennas

16.6.1 Early log-periodic antennas


The log-periodic principle was first established by Rumsey21 and Duhamel and
Isbell.9 It found its earliest and best practical application at HF, where the avail-
ability of decade bandwidths offered real savings in the number of antennas needed
and in the area of land required.
The HF antennas described by Duhamel and Berry8 were periodic structures in
which the dimensions of successive sections were increased in geometric progress-
ion (Fig. 16.45). The structures were fed at the vertex end, and energy travelled
along the structure until it reached a resonant portion. The resonant portion
radiated the energy, leaving only insignificant power in the larger sections beyond.
Provided that the periodic structure at the vertex end had dimensions small in com-
parison with the shortest operating wavelength, and provided that the large end was
700 High Frequency Antennas

large enough to radiate all the power at the longest operating wavelength, any
required bandwidth was theoretically possible.
The impedance of the early log-periodic antennas showed a marked periodic
variation, but generally remained within a VSWR of 2 : 1 ; thus they could be used
for transmission as well as reception. Free-space gains of around 6dB were readily
obtained, giving around lOdB with ground reflection reinforcement. Azimuth
beamwidths of around 60° were obtainable, with a backlobe perhaps 15dB below
the forward radiation. In comparison with the rhombic, the log-periodic was
inferior at the design optimum frequency, but superior when considered over a
wide band. The log-periodic did not need a matched termination and did not waste
any of the transmitted power.

balanced feed
Fig. 16.45 Log-periodic antenna

However, the early log-periodic had some disadvantages. From the designer's
point of view, it was not easy to control the antenna characteristics precisely
because the same periodic structure was acting both as transmission line and as
radiating element. A further major step forward was taken when the two functions
were separated, as they already were in the HAD and in the curtain arrays, resulting
in the logarithmic dipole antenna described in the following Section.

16.6.2 Logarithmic dipole antennas


The logarithmic dipole antenna was invented independently on both sides of the
Atlantic by Isbell11 and Kravis and Radford.14 It consists of a log-periodic version
of the HAD (Fig. 16.46), in which the transmission-line feeder is crossed between
each dipole to reverse the direction of fire. The logarithmic dipole antenna proved
easy to optimise, and its radiation patterns and impedance characteristics are sub-
stantially better than those of the early log-periodic.
The crossing of the feeder between each dipole element results in mutual cancel-
lation of backlobe components from the individual elements, and hence a very low
level of backlobe radiation, typically 25dB below main-lobe gain at HF and up to
35dB at VHF and UHF. Each element has smaller elements in front of it and larger
High Frequency Antennas 701

elements behind, and so tends to produce a directional primary pattern in the same
way as a Yagi antenna. This effect also contributes to the directivity and helps to
produce the very low backlobe and sidelobe level.

dipoles

feeder

Fig. 16.46 Log-periodic dipole antenna

The active elements, those at or just below resonance, are spaced at less than a
quarter-wave apart. The capacitive loading effects of the below-resonant elements
increase the effective spacing in the active region to a value closer to the quarter-
wave which theoretically provides the best impedance match. Furthermore, the
sharing of the power between several dipole elements in the active region and the
resultant tapered amplitude distribution improves both the impedance match and
the radiation pattern.
The theory of the logarithmic dipole antenna was first published by Carrel.7 His
calculated antenna characteristics showed good agreement with experimental results
obtained elsewhere.
For use at HF, the logarithmic dipole antenna is hung from a pair of masts with
its vertex end supported by low wooden poles so that the mathematical apex of
the dipole curtain lies in the ground plane (Fig. 16.47). This ensures that the
elevation patterns as well as the azimuth patterns remain independent of frequency.
For any given antenna and given angle of fire there is an optimum tilt angle. If the
masts are high, the ground reflection generates a shoulder lobe above the main
beam (Fig. 16.48). If too low, the main beam itself is at too high an elevation for
long-range working as in Fig. 16.49.
Vertically polarised logarithmic dipole antennas may also be used. Whereas the
azimuth beamwidth of a horizontally polarised antenna is typically 50° to 70°, that
of a vertically polarised antenna may be 100° or more; thus two antennas may be
combined in a vee formation to reduce the azimuth coverage (Fig. 16.50). If a dual
polarised antenna is required, it is possible to hang both horizontally and vertically
polarised antennas from the same pair of masts.
An interesting variant of the dual polarised logarithmic dipole antenna is shown
in Fig. 16.51. Two identical antennas are hung from a single mast and a single
common vertex pole, with inclinations of ± 45°. The two antennas are fed from a
hybrid transformer which gives either in-phase or antiphase excitation according
to which input terminals are used. The same antenna system can thus generate
702 High Frequency Antennas

Fig. 16.47 Horizontal log-periodic dipole antenna

30°tilt

Fig. 16.48 Log-periodic dipole antenna: Vertical radiation pattern

10°tilt

Fig. 16.49 Log-periodic dipole antenna: Vertical radiation pattern

feeders

Fig. 16.50 Vertical log-periodic dipole array


High Frequency Antenna* ~rno
either a low-elevation vertically polarised beam or a high-elevation horizontally
polarised beam (Fig. 16.52).
The logarithmic dipole antenna does not use all its aperture at any one time,
but merely the portion at or near resonance. Its aperture efficiency can therefore
be improved by increasing the bandwidth of the individual elements. During the
early 1960s a large number of designs were explored using folded dipoles, fan
dipoles and other elements of known good performance under isolated conditions.
The behaviour of these elements is modified when they are immersed in an array,
and some variants were more effective than others. The folded dipole gave a high
gain, but its impedance was poor and it could not be used for high-power trans-
mission.

feeders
Fig. 16.51 Dual-polarised log-periodic dipole array

horizontal
polarisation

vertical
polarisation

Fig. 16.52 Dual-polarised log-periodic dipole array: Vertical radiation patterns

One interesting variant was the broad dipole array invented by E. W. Woloszczuk
and described by Radford and Woloszczuk.20 In this antenna the dipoles were
broadened until they almost touched and were then replaced by rectangular wire
loops (Fig. 16.53). The action of the antenna was more complicated than the
simple explanation of element broad-banding would suggest, as was demonstrated
by current measurements on a full-size antenna, but the aperture efficiency was
some 2dB better than that of a simple dipole antenna. The improved impedance
match and very low sidelobes were additional advantages. However, the broad
dipole antenna was more difficult to rig than the simple dipole version, and it
required more wire and more insulators; thus it had no advantages in terms of cost-
effectiveness. As far as is known, the logarithmic broad dipole antenna has never
been investigated theoretically.
704 High Frequency Antennas
Loaded dipole elements have been used in logarithmic antennas. In rotatable
antennas used for communications with mobile units and other applications in
which mechanically steerable beams are desirable, inductive loading or capacitive
tip loading can be used to reduce the span of the longest dipoles (Fig. 16.54).
wire loop
dipoles

Dn ]DEL feeder

Fig. 16.53 Broad-dipole log-periodic antenna

capacitive tip loading series inductive loading

Fig. 16.54 Log-periodic antennas with loaded dipoles

Series capacitance loading can also be used. This has the opposite effect, increas-
ing the resonant length of the dipole but at the same time increasing the element
directivity in the plane of the dipoles. This increase is obtained without any accom-
panying increase in directivity in the other plane. By contrast, increasing the length
of a simple dipole array increases the directivity in both planes. The importance
of this will be seen in the next Section where arrays of logarithmic antennas are
considered.
In designing practical logarithmic antennas, it is first necessary to choose the
configuration appropriate to the specification. For short-range sky wave circuits,
requiring high elevation angle coverage, short horizontally-polarised antennas of
low taper ratio are appropriate. For longer ranges where higher gains and lower
angles of fire are needed a higher taper ratio is needed. The taper ratio t is defined as

t = (16.15)
High Frequency Antennas 705
where /„ and dn are the lengths and spacings of the nth dipole, respectively (see
Fig. 16.55).
It is also convenient to define the spacing ratio s:

s = |=- (16.16)
Lln
Many early designs used taper ratios such as 0.7943 (antilog — 0.1), 0.8913 (antilog
— 0.05) and 0.9441 (antilog —0.025), because this enabled lengths and spacings
to be read directly from antilogarithm tables.

Fig. 16.55 Geometry of log-periodic dipole antenna

10 • 100°

H-plane

E -plane
5- 50°
in .c
-D
T3
C
03 -


0-8 0-9 1-0

Fig. 16.56 Gain and beam width of log-periodic dipole antenna

Fig. 16.56 shows the variation of free-space antenna parameters with taper ratio
for a simple logarithmic dipole array, and Fig. 16.57 shows approximate optimum
ratios for given taper ratios. Ground reinforcement adds some 4dB rather than
706 High Frequency Antennas

the 6dB obtained with horizontal dipoles because the elevation pattern maxima
of the antenna and its ground image do not correspond.
In designing vertically-polarised logarithmic antennas, the usual problems of
ground conductivity arise, and extensive ground improvement is necessary unless
the antenna is on a marshy or coastal site. However, the design of the elements
follows the same rules as before.

0-2

0-8 t 0-9 10

Fig. 16.57 Log-periodic dipole antenna: Relationship between spacing and taper ratios

In a simple logarithmic dipole antenna the longest dipole should be half a wave-
length long at the lowest frequency of operation. It is usually convenient to short-
circuit the feed at this end; thus the last element is never driven but always acts as
a reflector. No terminating load is needed as little unradiated power remains at this
point. The shortest element should be resonant at about three times the highest
operating frequency of the antenna.
The transmission-line impedance is nominally 400-600 12, but since it is reduced
to 200-300 £1 by the capacitive loading of the dipoles on the vertex side of the
resonant region, it may be necessary to add a lumped capacitance at the point
where the feeder is connected to the antenna. This compensates for the discon-
tinuity in the capacitive loading; its value is best adjusted experimentally. With
correct compensation, an impedance match of 1.25 VSWR can be obtained with a
broad dipole configuration, or around 1.5 VSWR with simple dipoles.

16.6.3 Logarithmic antenna arrays


Logarithmic antennas provide less gain than optimised rhombics, and attempts to
increase the gain of logarithmic antennas by increasing antenna length show dimini-
shing returns. The mechanical design of very long logarithmic antennas becomes
difficult and expensive because of the need to keep all the dipoles properly tensioned.
The use of broadside arrays of logarithmic antennas provides an alternative solution
(Fig. 16.58). These arrays can then be used with electronic beam steering.
High Frequency Antennas 707

Broadside linear arrays of logarithmic antennas are limited in scan coverage by


the primary patterns of the array elements, reasonable maximum coverage being
±35° and ±55° for horizontally- and vertically-polarised antennas respectively.

to
phase
controlled
feeder system

Fig. 16.58 Logarithmic antenna array

These wide coverages are only obtained with short antennas of low taper ratio, and
result in high angles of fire in the horizontal case. If long-range low-angle coverage
patterns are needed, it may be necessary to resort to special logarithmic antenna
designs having parameters optimised for this application. Series capacitance-loaded
elements may be used to improve the elevation directivity of vertically-polarised
arrays without loss of azimuth scan coverage.
The array spacing can only be optimum at one frequency. At lower frequencies,
there is insufficient room for horizontally-polarised antennas while vertically-
polarised antennas are severely degraded by mutual effects. At higher frequencies,
the angle over which a unique beam is generated is reduced, until at frequencies
for which the spacing exceeds one wavelength two or more beams will always be
present. The amplitudes of the widely spaced grating lobes are, of course, reduced
by the shaping of the primary patterns, but grating lobes remain a potential source
of interference with other circuits in transmission and of unwanted noise pickup
and directional ambiguity in reception.

Fig. 16.59 Interleaved logarithmic antenna array

To avoid this problem, the idea of inserting small logarithmic antennas between
the larger antennas has been examined (Fig. 16.59), but the system becomes
unattractive when its detailed design and performance are considered. In particular,
there is an awkward frequency band just below the frequency at which the smaller
antennas can begin to contribute. Mutual effects cannot be neglected in this arrange-
ment; indeed they continue to degrade the array element performance out to
spacings of little over one wavelength.
708 High Frequency Antennas

An alternative scheme is shown in Fig. 16.60. This invokes the frequency-


independent principle and gives an increased gain without grating lobes as the
spacing of the active regions of the antennas in terms of wavelength remains constant.

direction
of fire

Fig. 16.60 Logarithmic antenna array

Fig. 16.61 Circular logarithmic antenna array

This system works well for arrays with few elements, but if a higher gain is required
the principle cannot be extended as the outer elements in the array would be firing
in the wrong direction. Fig. 16.61 shows a complete circular array using this prin-
ciple. It gives a full 360° coverage, but only combines those elements facing in the
right direction in any given beam position. Obviously a part-circular array can be
used if coverage is only required over a limited arc of scan.

16.7. Unorthodox HF antennas

16.7.1 HF horn antenna


The tapered aperture horn antenna (TAHA) was first described by Brueckmann
and Hagaman.6 It consists of a large horn made of wire mesh and suspended from a
large number of masts. For convenience of construction, the horn is of rhombic
section with a width about twice the height and a flattened floor (Fig. 16.62).
High Frequency Antennas 709

The horn is fed from a transmission line via a giant ridge waveguide section in
the throat. It is horizontally polarised.
The rhombic section results in some curvature of the field in the mouth of the
horn which tapers the excitation and so improves the sidelobe performance.

ridge waveguide wire mesh masts


feed in throat

side elevation section at mouth

Fig. 16.62 Tapered-aperture horn antenna

The bandwidth is limited by the matching of the feeder to about 3.5:1 for a
VSWR of 4. There is a considerable variation in the beamwidth over the frequency
range, but the sidelobes remain reasonably good, better than 20dB down at the
upper end of the band.
One unusual property of the TAHA is that the elevation angle of fire is less
frequency-dependent than that of conventional antennas having a fixed height.
This is because the phase centre height varies with frequency.
A practical full-size antenna covering 6.5-23 MHz measured some 257 m long,
154 m wide and 77 m high. Its gain varied from about 20dB at the lower end to
28dB at the upper end of the band.
The TAHA is, compared with a rhombic, expensive to construct and maintain,
and it has not proved as cost effective as the rhombic or the logarithmic antenna.

16.7.2 HF wire grid lens antenna


A two dimensional analogue Luneberg lens can be used to focus radio waves received
from any given bearing into a primary feed on the opposite side of the lens. A
practical HF antenna based on this principle was suggested by R. L. Tanner,24 using
a curious property of wire grids.
The velocity of propagation of a wave travelling between two wire mesh sur-
faces whose mesh size and spacing are small compared with the wavelength can be
varied between the velocity of light and l/\/2 of the velocity of light by varying
the ratio of the mesh size to the spacing between the grids. In Fig. 16.63 let the
grid spacing be 2a and the mesh be square with side b. When 2a is much greater
than b, propagation in any direction is much like that between two perfectly
conducting sheets. When 2a is much less than b, the structure becomes a network
of parallel-wire transmission lines. If propagation is at 45° to the side of the mesh,
then clearly the path is increased by a factor \fl. If propagation is parallel to one
side of the mesh, the grid wires at right angles to the direction of propagation act
as capacitive loading stubs and reduce the velocity by l/\/2.
If the propagation velocity is calculated precisely, as by Tanner and Andreasen,25
it is found that the wire-grid medium is not perfectly isotropic. The variations can
710 High Frequency Antennas

be reduced, at considerable additional expense, by using a hexagonal instead of a


square mesh. Other variations are introduced by the ratio between the structure
size and the wavelength, and also by the diameter of the wire used to construct
the mesh.

Fig. 16.63 HF wire grid lens antenna: Grid geometry

The equivalent refractive index (JL is given approximately by


(2m\ lb\
In sinh +
EL b I " \irrl
(16.17)
9 2na
In sinh
•nr b
where r is the radius of the wire. This formula needs to be modified if the mesh
size is less than about 100 times the wire diameter (see Tanner and Andreasen25
Part 2).
As it is not possible to obtain a refractive index much greater than about 1.3 in
practice by use of the wire-grid effect, it is not possible to construct a 2-dimensional
Luneberg with a focus on its circumference. A focal distance some 20% greater
than the radius must be used.

flare-

feed in throat
Fig. 16.64 HF wire grid lens antenna

From symmetry considerations, the lens can also be made by placing a single
grid over a perfectly conducting plane and making the spacing # instead of 2a.
Consider now the practical design of an HF antenna using the wire-grid lens
principle. The upper mesh can be suspended from a ring of non-conducting poles,
High Frequency Antennas 711

with spacers to separate it from the lower grid which is supported on stakes. A
conical flare surrounds the lens to shape the elevation pattern (Fig. 16.64). A
practical antenna covering 3-30 MHz had a diameter of 260 m and a flare height
of 27.5 m (see Sharp22).
The wire-grid lens antenna is expensive to construct and maintain, and can only
be justified if its multiple beam capability is fully exploited.

16.8 HF antennas for mobile applications

16.8.1 HF antennas for land vehicles


The design of antennas for vehicles must take account of certain factors which do
not arise in the design of fixed antenna systems. The vehicle itself is small in terms
of wavelengths, and if the equipment is to be able to operate when moving, the
vehicle will not have any satisfactory earth connection. The ground below the
vehicle may vary greatly in conductivity from one location to another. The vehicle
restricts the size and power input of the transmitter; thus the antenna is required
to radiate as efficiently as possible in order to provide acceptable performance,
yet the size of the antenna itself will inevitably be much less than that of an opti-
mum radiator.
Mobile communication systems are generally required to give highly reliable
service over short and medium rather than long ranges, often in difficult terrain.
In consequence, the lower end of the HF band is preferred, making antenna design
even more difficult. Either ground-wave or skywave propagation modes may be
needed.
To make operation independent of vehicle orientation, an omnidirectional
radiation pattern is required; thus some form of short monopole is the best practi-
cal solution. Rigid antennas are vulnerable to damage by trees and low bridges;
thus the monopole is usually made of springy steel or resin-bonded glass fibre, with
copper plating or a copper insert to improve conductivity. As the monopole must
be much less than a quarter wave high at the low end of the band some form of
reactive loading or compensation is needed to tune it to resonance and maximise
the radiation efficiency.
The vehicle is too small to form an efficient ground plane; thus the capacitance
between the vehicle and the ground appears in series with the impedance of the
antenna itself. The earth is lossy, and contributes an average of around 10 O to
the total input impedance of the antenna. For maximum efficiency the radiation
resistance must be the largest possible proportion of the sum of all the resistance,
including the copper, loading coil and earth losses.
Fig. 16.65 shows the approximate radiation resistance of a short unloaded
monopole. A 3 m whip is seen to have a radiation resistance of about 1 O at 5 MHz,
and so can be only about 9% efficient even if copper losses and tuning losses can be
neglected. The resistance of the base loading inductance used to tune the antenna
to resonance further reduces the efficiency, but note that the losses also assist the
712 High Frequency Antennas

matching by raising the input impedance from 1 £1 to perhaps 15-20 £2, which is
easily transformed to 50 £2 for feeder connection. The calculation of loading coil
inductance follows the basic theory of Chapter 1 (see also Chapter 15 and Belrose2).

E
JC
O

0 004 0-08 0-12


height (A)
Fig. 16.65 Radiation resistance of a short monopole

An elevated loading coil can be used to improve the antenna current distribution
and thus raise the radiation resistance. The optimum coil position varies with the
earth loss, which is itself variable and unpredictable, but halfway up the whip is
a good practical compromise. The elevated loading coil gives a small but worthwhile
improvement in efficiency as compared with the base loading coil position, but
there are two attendant disadvantages:
(i) The inductance is no longer easily accessible for tuning when the frequency is
changed.
(ii) The loading inductance complicates the mechanical design of the whip, requir-
ing a stronger lower portion to support its weight and increasing its vulner-
ability to accidental damage.
If a wide tuning capability is required, the use of an elevated loading coil is thus
hardly worthwhile. If fixed-frequency or narrow-band operation is sufficient and
if the operational environment is not too difficult, a fixed elevated loading coil
may be used. Note that even if the coil inductance is calculated to make the input
impedance purely resistive on any one site, additional tuning reactance will still be
needed to compensate for changes in site characteristics.
The shorter the whip and the greater the inductance needed to tune it to reso-
nance, the narrower will be the instantaneous bandwidth. The instantaneous band-
width of a typical whip antenna is only around 1%, which is adequate for voice
or teleprinter operation but leaves little scope for frequency changing without
High Frequency A n tennas 713

retiming. If frequent changes are necessary an automatic tuning system becomes


justifiable.

16.8.2 HF an tennas for aircraft


Antennas for aircraft and helicopters are perhaps the most difficult of all HF
antennas to design. The aircraft itself is a difficult environment, with its wide
variations in temperature according to altitude and speed, its vibration, and its
severe mechanical stresses produced by aerodynamic loading. Equally difficult
are the constraints imposed by the aircraft designer, who demands that the antenna
should neither significantly increase the drag nor reduce the strength of the structure.
Like the vehicle, the aircraft may be small in terms of wavelengths, particularly
at the lower end of the band which is most useful for air-ground communications.
A small aircraft or helicopter may always be below resonant size, while a larger air-
craft may have several different modes of resonance within the HF band limits.
In either case, the HF antenna is not an antenna in isolation but a coupling device
which enables the transmitter to energise the aircraft structure.
The first step in solving the problem is to discover what resonances occur, and to
select a feed point which will give the best excitation of the wanted modes through
the working part of the band. At the same time, the feed point must minimise the
variation in impedance so that the losses in the tuning unit can be kept to a mini-
mum. One obvious technique for analysing the possible resonant modes is wire-
grid modelling. An alternative technique is scale modelling, and to avoid the errors
produced by cables and interactions between the model and the measuring equip-
ment a complementary 'Babinef model may be used (see Bolljohn and Granger4).
In Fig. 16.66 the aircraft is represented by a slot in a conducting sheet. Since the
aircraft is symmetrical, a half slot and an image plane is sufficient. The slot is fed
by means of a coaxial cable on the reverse side of the ground plane. The impedance
of the aircraft Za is given by

I. - ()

where Zs is the equivalent slot impedance. This method cannot be perfectly accu-
rate as it uses only a 2-dimensional projection of the aircraft fuselage, but it gives
the designer an initial insight into which modes are present. On a large aircraft,
some part of the structure such as the wing or the tail may become quarter-wave
resonant, and will present a low impedance and so draw most of the power. It will
then dominate the radiation pattern, and may result in little or no power being
radiated in some directions.
The earliest HF aircraft antenna was the 'clothes-line', a wire antenna supported
by a short mast at the forward end and the tail fin at the aft end. The clothes line
actually forms a slot of sub resonant size, and external reactance is needed to tune
it to resonance. The clothes line was satisfactory for slow piston-engined aircraft,
but its drag was not acceptable on high-performance aircraft.
To meet the needs of the aerodynamicist, completely suppressed antennas were
714 High Frequency Antennas

devised, that is, antennas buried in the aircraft structure. Examples include notches
(Fig. 16.67), tail and wing caps insulated from the main structure, slots, and con-
ducting rails buried in insulated fairings. Small probes having minimal drag have
also been used on large aircraft.

coaxial feed
image plane

conducting
sheet

Fig. 16.66 Complementary 'Babinet' model of an aircraft

notch in /
fin fairing^/

Fig. 16.67 Notch antenna

Of these various possibilities the notch first proposed by Johnson12 is the most
efficient. Unfortunately, the ideal electrical position for the notch is usually a wing
root or at the base of the tail fin, both highly-stressed regions where the inter-
ruption of a spar by an insulated joint is mechanically undesirable. As a result, the
notch cannot be made as large as the antenna designer would like, and its impe-
dance is normally inductive. Thus, it has to be tuned to resonance by a capacitative
tuner.
The impedance of the tuned notch may be very different from the 50 £2 impe-
dance of the feeder cable; thus a matching transformer is also required. Since the
slot impedance varies with frequency, a simple transformer is not enough and an
adjustable reactive matching network must be used. The tuning and matching func-
tions may be separated, or they may be combined in a single unit. Fig. 16.68 shows
the arrangement of a typical notch and its associated tuning and matching equip-
ment.
For efficient operation, the notch must be coupled as tightly as possible into the
High Frequency A ntennas 715

currents flowing on the aircraft surface. In practice, even when the notch is opti-
mally sited at a high-current point like a wing root its radiation resistance is only
around 0.1 ^2. Low-loss reactances must be used for tuning and matching, and the
tuned notch may have a useful bandwidth of as little as 10 kHz.

Fig. 16.68 Notch antenna and tuning unit

Since several frequencies may have to be used to maintain communication under


different ionospheric conditions and at different ranges, the notch may need to be
re tuned in flight. A remotely controlled or automatic tuner is required. Tuning may
be performed by mechanical variation of reactances during the first few seconds
of transmission until minimum reflection is obtained at the feed point, or more
conveniently by electronically switched reactances set up from stored data just
before transmission commences.
Cap and slot antennas, like the notch, present large reactances and low radiation
resistances and require similar tuning and matching units.
Two other antennas deserve a mention. The trailing wire can be used on low-
performance aircraft when well clear of the ground, but it has to be wound in
before landing. It has an appreciable drag. Tuning can be carried out by adjustable
reactances at the feed point or by varying the amount of wire extended.
The remaining antenna is the loaded whip. This cannot be used on fast aircraft,
but it is useful on helicopters where the structure itself is usually too small to be
energised by notches or other means. The design of the whip follows the principles
discussed earlier.
One further problem which arises on aircraft is that of operating on two or more
circuits simultaneously, particularly when, for ionospheric reasons, the frequencies
cannot be widely separated. There is as yet no really satisfactory solution to this
problem, but in principle two different feed points can be used, selected to give
minimum mutual coupling, and the tuning and matching units can be arranged to
minimise pick-up of each other's transmissions.
Substantial transmitted powers may have to be used to overcome the losses in
aircraft antenna systems. This means that high currents and high voltages will be
present, and these must be allowed for in the design, together with the effects of
reduced atmospheric pressure on voltage breakdown.
716 High Frequency Antennas

The low antenna efficiency plus the possibility of noise generated by other
electronic equipment and static discharges means that the receiving signal/noise
problems cannot be forgotten. Losses in reception must be kept low, all inter-
ference from other equipments must be carefully suppressed, and antennas must
be chosen to minimise static noise. The notch is one of the best antennas in this
respect and this is a further reason for its popularity.
The provision of efficient HF antennas in aircraft will become more and more
difficult as the constraints of aerodynamics, space and weight become more severe.
However, the availability of satellite communications will relegate HF to emergency
back-up status, whereas until recently HF provided the only means by which an
aircraft in mid-ocean could communicate with land.

16.8.3 HF antennas for ships


Ships need omnidirectional antennas in order to be able to communicate with
stations on any bearing. Since the sea has a relatively high conductivity, simple
vertically polarised radiators meet the need adequately, enabling sky-wave or
ground-wave propagation to be used according to the range of the distant station.
Marine antennas must be designed to withstand the severe environment, which
includes the physical stresses of wind and ice loading and the possibility of damage
by heavy seas, together with chemical attack by salt spray and funnel gases. They
must also be able to operate in the electrical environment of the ship, picking
up minimum interference from other equipment and also causing minimum inter-
ference to others. In particular, it may be necessary to operate two or more HF
equipments simultaneously without mutual interference.
Early marine HF antennas were substantial L or T-monopoles slung between
masts, usually with several horizontal wires to maximise the top capacitance load-
ing. Recent practice has been to eliminate large HF antennas and install whip
antennas instead. These can be placed close to the transmitters, minimising the
length of transmission line. Since two or more whips can be carried, separated by
several tens of metres and with some superstructure between, the mutual isolation
of antennas is improved. On warships where many other electronic equipments
may be carried, including some needing unobstructed sites, the advantages of
relegating the HF antennas to secondary positions are particularly significant.
The electrical design of whip antennas for marine use is the same as that for
vehicle antennas as described in the previous Section, except that rather longer
whips can be used with a corresponding improvement in efficiency.
In siting the whip, it is necessary to take into account the effects of masts,
funnels and other obstructions. The best antenna position can be found by scale
modelling. Given a sufficiently detailed model, mathematical optimisation is also
possible, and it is likely that, as computing powers increase, theoretical methods
will gradually replace scale modelling. Currently, both methods are applicable, the
choice depending upon the complexity of the problem.
An unusual alternative to the vertical monopole and whip class of antennas is
the notch. It is possible to excite a funnel or other suitably shaped part of the
High Frequency A n tennas 717
structure by cutting a notch near the base and incorporating a remotely controlled
tuning device, as described in the previous Section on aircraft antennas. While this
gives a cleaner appearance, it is less convenient and more expensive to install than
a whip, and is unlikely to replace the whip in normal marine applications.
Bearing in mind the steel decks and wet surfaces, particular care is necessary in
designing and installing HF antennas so that they cannot harm personnel.

16.9 HF antenna design

16.9.1 Theoretical design techniques


Of the various theoretical methods available for the HF band, the use of thin wire
grid models is generally the most convenient. Antennas like rhombics and logarith-
mic dipole arrays can be analysed directly using available computing capabilities,
regarding each wire as sufficiently thin that the current flow is at the centre and
assuming perfect conductivity.
Antennas on vehicles, aircraft and other finite conducting bodies can be analysed
by wire-grid modelling. The grid spacing should be small, preferably not more than
0.05A, but since this corresponds to 1 m at 15 MHz and since most bodies of
interest are only a few metres in extent, the models need not be unduly complicated.
Whichever computing technique is adopted, it should be rembered that the
apparent precision of a computed solution may not be justifiable in view of the
uncertainty of the input data. At HF, antenna performance is not really separable
from that of the ionosphere and the ground, and a near-optimum design will be
operationally indistinguishable from a (momentarily) optimum design.

16.9.2 Scale-modelling techniques


Recent advances in theoretical modelling techniques have reduced but not elimin-
ated the need for scale modelling of new HF antennas. Scale modelling enables
second-order practical effects to be taken into account, and provides the insight
into antenna operation which inspires improvement and innovation. Scale model-
ling also provides one of the best ways of explaining and demonstrating a new
design to potential users.
The theory of scale modelling has been established by Sinclair.23 He shows that,
in a scale model, wavelengths, capacitances and inductances are reduced in pro-
portion to the linear dimensions, while conductivities are increased in the same
ratio. Other parameters such as gains, impedances, beamwidths, sidelobe levels,
dielectric constants and permeabilities remain unchanged in a perfect model. If
the conductivity of the full-size antenna is good, the scaled increase in conduc-
tivity makes little practical difference and copper wire may safely be used to model
copper-cadmium or steel-cored aluminium conductors. Poor conductors like real
earth are less easily scaled as it is necessary to increase the conductivity while main-
taining the dielectric constant unchanged. Scale models are often tested over per-
fectly conducting ground planes to avoid this difficulty, the measured performance
being adjusted empirically to compensate for ground losses.
718 High Frequency Antennas

At least two classes of scale model may be used. Small-scale models having scale
factors of order 1/100 are suitable for initial investigations of new designs and for
measurement of radiation patterns. Large-scale models having scale factors of 1/2
to 1/5 are suitable for impedance measurements and for initial trials. Since their
construction can be similar to that of the full-size antenna they also provide useful
experience for the mechanical designers and riggers.
In choosing the best scale for any model, the size of the antenna test site, the
available equipment, the constructional facilities and the nature and accuracy of
the information required must be taken into account. True radiation patterns can
only be measured in the far field, which is usually taken to be the region beyond
the critical range R given by
ID2
R = — (16.19)

where D is the antenna aperture and X the wavelength. For precise measurement of
very low sidelobes, a somewhat longer range should be used. The site thus deter-
mines the lowest frequency which can be used for antenna pattern measurements.
The highest frequency which can be used is determined by the amount of detail
to be reproduced in the model and the frequency and bandwidth of the measuring
equipment available.
Impedances may be measured accurately in coaxial lines over a frequency range
extending from HF to microwaves, but accurate measurement in balanced open
lines becomes difficult above the VHF band. Balance-to-unbalance transformers
covering wide bands are not easy to construct with sufficient accuracy; thus for
precise impedance measurements on balanced antennas large-scale models should
be used.
The construction of small-scale models cannot easily reproduce all the features
of full-size antennas. Wire cannot be precisely scaled, or it becomes extremely
fragile and its ohmic resistance is too high. It is better to use copper wire of suf-
ficient diameter to stand up to normal careful handling, and recognise that impe-
dances and losses may not be perfectly represented in the model. Again, rigging
cannot be precisely scaled in a 1/100 model but unless there is some question
of obscuration or reradiation by the rigging, the radiation patterns will not be much
affected. Satisfactory models of dipole arrays, curtains and logarithmic antennas can
be made by supporting the radiating wires on light perspex frameworks.
Models of aircraft, vehicles and ships can be used to optimise antenna positions
and predict radiation patterns when the structures are too complex for theoretical
modelling. However, the models must be reasonably detailed, and must include
any other antennas or other structures which are likely to be added. With careful
modelling, approximate prediction of antenna impedance becomes possible. The
cost of a detailed model is usually amply justified by the time saved when the full-
size antenna is installed.
High Frequency Antennas 719

16.9.3 HF antenna performance measurement


While theoretical and scale-modelling techniques permit precise estimates of antenna
gain, sidelobe level and match to be obtained, the accuracy of these estimates
depends on the accuracy with which the model reproduces the characteristics of
the full-size antenna. The errors made in assessing the qualities of the site may be
considerably greater than any approximations made in the modelling. There may
therefore be a need to measure the installed performance of the antenna.
Because of the difficulty of estimating absolute gain, measurements of gain are
usually based on comparison with a known simple antenna such as a dipole on an
adjacent site, mounted at the same effective height so as to have as nearly as possible
the same elevation radiation patterns. Obviously, the comparison antenna must
have the same polarisation. It must be far enough away to eliminate mutual coupl-
ing or obscuration, yet near enough to have the same site characteristics.
Known signals can be used for the measurements if sufficient time is available
to collect enough results to reach a statistically meaningful result. A precision vari-
able attenuator is used to equate the signal levels received from the two antennas,
the receiver being switched several times between the two until a best estimate is
obtained. The actual instantaneous levels may well be varying due to multipath
interference at the two antenna positions. For antennas with a low angle of fire,
long-range stations should be chosen and a frequency band close to MUF employed
to eliminate high-angle signals. For high-angle antennas local stations just beyond
the skip distance should be chosen.
The measured gains with respect to the comparison antenna are then plotted on
a theoretical diagram of the comparison antenna. A scatter of several decibels may
be found, but by drawing the best curve through the plots the antenna radiation
pattern can be obtained. This method is particularly useful for demonstrating the
sidelobe and backlobe protection of antennas.
When a quick result is needed, an aircraft or helicopter can be flown around the
antenna with a small transmitter. This method is expensive, and is not particularly
accurate as the polarisation and radiation pattern of the airborne antenna may not
be accurately known. However, the precise direction and elevation of the source
may be measured by a combination of electronic and visual tracking techniques,
whereas with signals from known distant stations, the angles are uncertain because
of uncertainty about the ionospheric conditions over the signal path.
Measurements of antenna match are easily carried out at the transmitter, but
any feeder imperfections are then included in the result. Measurements of antenna
match can also be carried out directly at the feed point, but great care is then
needed to shield the impedance bridge from stray pickup since it is in the radiated
field. The presence of the operator in the near field may also affect the result, and
unless a remote control is available the operator may find it necessary to lay flat
on the ground.
Measurements of antenna efficiency are exceedingly difficult to make. Some
estimate of relative efficiency may be made by comparing the measured gains with
those calculated from measured or theoretical radiation patterns of the antenna
under test and the comparison antenna.
720 High Frequency Antennas

16.9.4 Prac tical HF an tenna design


Most HF antennas are of wire construction. The radiating wires should be tensioned
to keep them straight and to ensure that any movement due to wind or ice loading
does not distort the antenna beyond the electrically acceptable tolerances. Tensions
may vary from a few tens of kilograms to several hundred kilograms according to
the length and permissible movement of the wire. Simple antennas like rhombics
have their radiating wires tensioned directly between the supporting masts, with
only short support wires between the insulators and the masthead. More elaborate
antennas like logarithmic dipoles have their radiators tensioned between heavy
catenary supports, as in Fig. 16.69, the tension tn in the nth dipole being given
approximately by the formula

tn = — tan0 n ) (16.20)
where <j)n is the inclination of the catenary between the (n — l)th and nth dipole
and T is the tension of the section whose inclination is zero. This formula assumes
that the weight of the catenary itself is negligible compared with the dipole tensions.

v t

Fig. 16.69 Tension diagram for catenary support wires

The catenary tensions in large antennas may be several thousand kilograms. Mech-
anical design involves finding a compromise between the catenary tension and the
permissible curvature, since the lower the catenary tension the greater must be the
total span of the catenaries at the end of the array. In the case of a vertical array,
the lower the catenary tension the greater must be the height of the masts above
the top of the radiating elements.
The tensions in an ice-loaded antenna under wind-loaded conditions will be
much in excess of the normal working loads. A wire catenary which is only slightly
elastic suffers a greater excess load than a synthetic rope catenary which stretches
under load and so eases the dipole tensions. However, synthetic ropes are less stable
than wires and are generally only used on smaller antennas. The antenna should
always be designed to fail safe, so that under intolerable loads it is an antenna wire
or an insulator which gives way and not a mast or a main guy wire. For antenna
design in the United Kindgom, an antenna designed to withstand a 15 m/s wind
with a 5 mm radius ice loading represents a reasonable compromise between surviv-
ability and cost.
High Frequency Antennas 721

If conductivity were the only consideration, wire antennas would be made of


pure copper or aluminium. Unfortunately copper work-hardens, becomes brittle
and fractures, while pure aluminium is relatively weak in tension. Until a few years
ago, cadmium-copper alloy was the preferred material as it does not work-harden
and its conductivity, although inferior to that of copper, is still reasonably good.
Wires of around 2 mm diameter were usually chosen. More recently, composite
wires haveing a steel core for strength and an outer sleeve of aluminium for con-
ductivity have taken the place of copper alloys.
Non-radiating supports may be of synthetic rope or of galvanised steel broken
into short lengths by insulators. Break-up insulators are also inserted in guy wires
used to support masts, and may even be necessary in the masts themselves if they
obstruct the radiated fields. However, it is more usual to employ wooden poles
where height requirements are modest, as for example at the vertex end of log-
arithmic antennas. Where possible, masts should be orthogonal to the radiated
fields, as in the HAD, or behind the radiating wires so as to act as additional reflec-
tors, as in logarithmic antennas.
Insulators are usually of ceramic or polypropylene. If possible, they are designed
so that they are in compression rather than in tension, and so that failure of the
insulator does not result in collapse of the antenna. The 'egg' insulator fulfils these
requirements but has only limited power-handling capability. For high-power trans-
mission more elaborate insulators are needed to avoid voltage breakdown in un-
favourable weather conditions.
Masts are preferred to free-standing towers where space is available for the guy
wires, since masts are both cheaper to manufacture and easier to install. Small masts
may be of tubular steel, but triangular lattice masts are usually preferred. They
are built up in standard sections for heights of 30-60 m. Small masts may be assem-
bled in a horizontal position and then hoisted upright using a short auxiliary mast.
Larger masts are usually built up one section at a time using a jury pole to support
each succeeding section until it is safely bolted in place and its guys attached.
The erection of masts and rigging of antennas is highly skilled work and should
not be undertaken by inexperienced personnel. Rigorous safety procedures must
be followed at all times. This warning may seem unnecessary in a chapter on
antenna design, but nevertheless it is not unknown for theoretically qualified
engineers to be injured when attempting practical rigging tasks.
Masts and guys are usually anchored to massive concrete blocks cast in situ in
previously surveyed positions. For small masts and temporary or transportable
installations, steel pegs or steel screws may be used as anchors. Special anchoring
arrangements may be needed on rocky sites or soft beaches.
In comparison with wire antennas, rigid monopoles and whips raise few con-
structional problems. Rigid receiving monopoles are usually of tubular construc-
tion, enabling small transformers to be placed in the hollow centre close to the feed
point, with coaxial feeder running down the centre to the receiver.
When antennas are mounted on vehicles or used to excite aircraft, currents flow
on the surface of the vehicle or aircraft. It is important that any joints in the
722 High Frequency Antennas

structure should be electrically bonded, and that imperfect joints which could cause
crossmodulation or generate noise should be eliminated. It is also important that
any ignition systems or other electrical equipment on the vehicle should be care-
fully screened and decoupled.
Transmission lines for HF antennas must have a good impedance match, low
losses and adequate power-handling capability. Older transmitting antennas usually
had open-wire balanced feeders of impedance 600 £2 (two-wire) or 300 Q (four-
wire). Attenuation in open feeders is low, typically 1-2 dB/km, and power handling
is high, tens or even hundreds of kilowatts with suitable spacing and insulation.
Open feeders require careful design at bends or twists, as any unbalance can
result in leakage of energy as well as degradation of the impedance match. If open
feeders are allowed to pass too close to each other, crosstalk can occur. Open
feeders are subject to wind and ice loading, and their mechanical design must take
these factors inco account.
Coaxial feeders are now preferred for most applications. They overcome leakage
and crosstalk problems and are immune from ice and wind loading, but their
initial installation costs are higher. The attenuation of large-diameter concentric
lines is comparable with that of open-wire feeders, and power levels of up to 100
kW can be handled by coaxial lines of around 10 cm diameter. Impedances of
50 £2 are usual for transmission, this being the best compromise between low
attenuation and high-power-handling requirements. For reception 75 £2 is some-
times used on account of its marginally lower attentuation, but recent designs tend
to standardise on 50 £2 for both transmission and reception.
High-power concentric line feeders must be sealed and are usually filled with
dry air or nitrogen at slightly above atmospheric pressure to prevent ingress of
moisture.
On large HF communication sites there may be a switching centre which enables
any transmitter to be coupled to any antenna.
Complaints from users about antenna performance can often be traced to faults
in feeders and switching centres. In particular, crosstalk and leakage can be inter-
preted as sidelobe pickup. Broken or out-of-balance feeders can radiate and receive
sufficient power to maintain communication under good ionospheric conditions,
and this can lead to complaints of low signal strength rather than absence of signals.

16.9.5 HF antenna siting


The choice of a site for an HF antenna system is seldom left to the antenna designer,
since political, economic, logistic and other factors usually predominate, but the
designer can improve the antenna cost-effectiveness by optimising the use made of
whatever site he is given.
The ideal site should have adequate space for all the antennas to be deployed
and should allow reserve space for future growth. There should be open country,
or better, open sea, in the directions of fire, and there should be no obstructions
such as overhead power lines in these directions. A remote site is preferable, giving
freedom from man-made interference and freedom from interference with other
High Frequency Antennas 723

services. However, there must be reasonable site access for construction, and reason-
able domestic and other facilities for operators.
The site must be protected from vandalism and pilfering, copper being par-
ticularly attractive to thieves. It is also necessary to protect innocent members of
the public from high voltages and from intense radiation fields.
Having found a site, the individual antennas must be carefully placed to mini-
mise mutual obscuration and interference. It should be noted that wideband
antennas can cause more obscuration problems than highly resonant antennas, since
they are active over a wider band. If the site is an existing station with antennas
already present, the possibility of resiting some of the existing antennas should be
considered. In determining which positions give least mutual disturbance, a scale-
modelling or theoretical-modelling exercise may be justifiable.
The risk of accidental ignition of inflammable vapour at nearby petrochemical
plants and gas terminals by high-power HF transmitters should also be checked
when any new site is assessed.
Environmental acceptability can be improved by using low ground, where con-
ductivity may also be better, and by using slender masts rather than heavy towers.
In undulating country, antennas on the slopes of hills do not break the skyline and
are far less noticeable. A site which slopes gently downwards in the direction
of fire also improves the low elevation performance of the antenna.
The antenna engineer may protest that environmental considerations are outside
his terms of reference. However, he should remember that planning authorities
and the general public at large understand little of electromagnetic theory, and that
he is better able to optimise the environmental impact of his work than anyone
else. The acceptance of this responsibility is the inescapable duty of the professional
antenna engineer.

16.10 References

1 ANDERSEN, J. B.: 'Influence of surroundings on vertically polarised log-periodic anten-


nas', Teletechnik, 9, 1965, p. 33
2 BELROSE, J. S.: 'Short antennas for mobile operation', QST, Sept. 1953, p. 30
3 BEVERAGE, H. H., RICE, C. W., and KELLOGG, E. W.: 'The wave antenna - a new type
of highly directive antenna', AIEE Trans., 42,1923, p. 215
4 BOLLJAHN, J. T., and GRANGER, J. V. N.: 'The use of complementary slots in aircraft
antenna impedance measurements', Proc. IRE, 39, 1951, p. 1445
5 BRUCE, E.: 'Developments in short wave directive antennas', Proc. IRE, 19, 1931, p. 1406
6 BRUECKMANN, H., and HAGAMAN, B. G.: 'Horn antennas for high frequency long
range communication', IRE Trans., AP8,1960, p. 523
7 CARREL, R.: 'The design of log-periodic dipole antennas', IRE Intemat. Conv. Rec, 1,
1961, p. 61
8 DUHAMEL, R. H., and BERRY, D. G.: 'A new concept in high frequency antenna design',
IRE Nat. Conv. Rec, 1, 1959, p. 42
9 DUHAMEL, R. H., and ISBELL, D. E.: 'Broadband logarithmically periodic antenna
structures', IRE Nat. Conv. Rec, 1,1957, p. 119
724 High Frequency Antennas

10 HANSEN, W. W., and WOODYARD, J. R.: 'A new principle in directional antenna design',
Proc.IRE, 26,1938, p. 333
11 ISBELL, D. E.: 'Log periodic dipole arrays', IRE Trans., AP8,1960, p. 260 (also Univ. of
Illinois Ant. Lab. Tech. Report 39, June 1959).
12 JOHNSON, W. A.: "The notch aerial and some applications to aircraft radio installations',
Proc. IEE, 102B, 1955, p. 211
13 KNIGHT, P.: 'The selection of the optimum curtain array for a high-frequency broadcast-
ing service', Proc. IEE, 109, Pt. B, 1962, p. 91
14 KRAVIS, A., and RADFORD, M. F.: British Patent No. 884889. 1959
15 LADNER, A. W.: 'The series phase array', Marconi Rev., 67,1937, p. 1
16 MASON, H. P.: 'Some factors influencing the design of broad band HF monopole aerials',
IEE HF comm. Conv. ED4,1963, p. 114
17 MORRIS, D. W., MITCHELL, G., MAY, E. J. P., HUGHS, C. J., and DALGLEISH, D. I.:
'An experimental multiple-direction universally steerable aerial system for HF reception',
IEE HF Comm. Conv. ED4,1963, p. 121
18 MOXON, L. A.: 'High frequency antennae and propagation modes in relation to the
amateur service', IEE Ant. and Prop. Conf., Pt. 1, 1978, p. 83
19 NORMAN, F. J., and WARD, J. F.: 'Rhombic aerials, design charts for high frequencies',
Electronic & Radio Engr., 34,1957, p. 398
20 RADFORD, M. F., and WOLOSZCZUK, E. W.: 'New HF logarithmic aerials', IEE HF
Comm. Conv. ED4,1963, p. 141
21 RUMSEY, V. H.: 'Frequency-independent antennas' IRE Nat. Conv. Rec. 1, 1957, p. 114
22 SHARP, E. D.: 'Electromagnetic theory of wire grid lens HF antenna', IRE Trans., AP13,
1965,p.703
23 SINCLAIR, G.: 'Theory of models of electromagnetic systems', Proc. IRE, 36, 1948,
p.1364
24 TANNER, R. L.: 'New ideas in HF antennas', IRE PGAP Meeting, Stanford Univ., Jan.
1960
25 TANNER, R. L., and ANDREASEN, M. G.: 'A wire grid lens of wide application', IRE
Trans., AP10,1962, p. 408
26 WAIT, J. R., and WATERS, L. C: 'Influence of a sector ground screen on the field of a
vertical antenna', NBS Sci. Report no. 22. Oct. 1962
27 WATT-CARTER, D. E., and YOUNG, S. G.: 'A survey of aerials and aerial distribution
techniques in the HF fixed service', IEE HF Comm. Conv., ED4, 1963, p. 67
28 BENNINGTON, T. W.: 'Short-wave radio and the ionosphere' Iliffe and Sons Ltd, London,
1950.
29 DA VIES, K.: 'Ionospheric radio propagation' National Bureau of Standards Monograph 80,
Washington, 1965
Chapter 17

VHF and UHF antennas


R. A. Burberry

17.1 Introduction

17.1.1 Definition of frequency range


VHF (very high frequency) is generally defined as including frequencies from 30 to
300MHz (wavelengths 10m to 1 m). In a number of applications, the term VHF is
associated with limited portions of the complete range, where the principal appli-
cations are:
Military communications: 30-76 MHz
Sound radio broadcasting: 87.5-108 MHz
Aeronautical communications and radio navigation: 108-136 MHz
Marine communications: 158-162MHz
Television broadcasting: 174-230 MHz
Similarly, whilst the term UHF (ultra high frequency) is generally understood to
mean 300 to 3000 MHz (wavelengths 1 m to 0.1 m), in the field of military com-
munications the term covers the range 225 to 400 MHz. The frequency range 470
to 862MHz is used for television broadcasting. In this Chapter, UHF will be limited
to an upper frequency of 1 GHz (0.3 m).

17.1.2 Applications
The range of wavelength permits the use of antennas whose dimensions are of the
order of a half or one wavelength. The antennas are physically convenient and
mechanically simple, and are one of the reasons for the widespread use of these
frequencies.
The predominantly line-of-sight propagation characteristics of VHF and UHF
allow repeated use of the same frequencies within comparatively small distances,
and so make these frequencies well suited for area coverage systems. At the same
time, it is practical to construct antennas having moderate gain and directivity so
that these bands are also used for fixed point-to-point links. Typically, a complete
communication or broadcasting network, in an area where telephone lines are
726 VHF and UHF antennas

impractical, might consist of a mesh of point-to-point links, at UHF or higher fre-


quencies, each node having an associated area coverage system at VHF.
Most practical applications favour the use of simple but efficient antenna sys-
tems. Most short-range mobile communications and navigation systems are at VHF
or UHF, lower frequencies only being preferred in terrain which offers poor line-of-
sight characteristics.
Some common applications for VHF and UHF are listed below:
Broadcast (radio and television)
Point-to-point communication
Mobile communication
Navigation
Direction finding

17.1.3 Classes of antenna


Antennas which are well suited to the VHF-UHF bands can be considered in broad
categories based upon their principal frequency or radiation characteristics as
follows:
Omnidirectional (in one plane)
Directional
Broadband
Multiband
Steered beam
Circularly polarised
The antennas described in this Chapter will be categorised by electrical type
rather than by application, as it will be seen that the same types can be used in a
number of applications. Additionally, a section will be devoted to the problems of
installing a number of antennas in close proximity, whether it be on a common
mast or on a vehicle.

17.2 Omnidirectional antennas

The term 'omnidirectional' is used rather loosely in practice to indicate uniform or


near-uniform coverage in the horizontal plane. In the simplest form this require-
ment may be met by a small loop or a vertical dipole, but vertical arrays of such
elements may be employed to provide additional gain or to modify the radiation
pattern in the elevation plane while maintaining omnidirectional coverage in the
horizontal plane. In some cases, the support structure may modify the radiation
characteristics sufficiently to necessitate the use of radiating elements arranged in a
circle rather than a single central radiator.
VHF and UHF antennas 727
17.2A Single elemen ts

Vertically polarised dipoles: The axial symmetry of a vertical dipole makes it a


natural choice as a single element. Practical considerations in the feeding and sup-
port of such an antenna often make the centre-fed symmetrical dipole less con-
venient than would at first appear. Consider, for example, the arrangement of Fig.
17.1 a. The practical antenna will consist of the dipole, a balun (balance to unbalance
transformer), a coaxial cable and a support structure. The latter will have some
effect on impedance and radiation pattern and also on coupling to other antennas
on the same structure. These factors are examined in Sections 17.2.3 and 17.8. In
many instances, therefore, it will be preferable to use an axially-fed dipole rather
than one fed symmetrically in the conventional manner.

A/2
H

A/4

Fig. 17.1a Basic A/2 dipole with balun


b Same dipole with sleeve

Whilst the simple construction of Fig. 17.1a will be suitable for temporary
measurement purposes, the dipole for long-term exposure is more likely to take the
form of Fig. 17.1Z?. The balun and centre of the dipole are encased in metal, thus
providing mechanical support at the base of the antenna, protection of the terminals
from the weather, and the facility for supporting the antenna by the balun. The
antenna thus formed is termed a sleeve dipole. For overall lengths of the order of a
halfwavelength the vertical radiation pattern can be regarded as identical with that
of a plain cylindrical dipole.
The equivalent circuit for the impedance of a sleeve dipole is shown in Fig. 17.2
in which ZA represents the impedance of each half of the dipole at the top of the
sleeve, XF is the shunting reactance due to the discontinuity at this point (normally
capacitive) and Z o and B are characteristics of the coaxial line formed by the antenna
element and the sleeve. When the length of each antenna element is approximately
a quarter wavelength the resonant impedance at A is RA =R0/cos2d, where ^ 0 is
the basic impedance of the antenna element alone. Fig. 17.3 shows measured
impedances at the top of the sleeve for a range of sleeve monopoles. In general, the
sleeve increases the radiation resistance and the capacitive reactance.
728 VHF and UHF antennas

The dipole in Fig. 17.1 employs a balun to permit feeding from coaxial line.
Below about 150MHz, quarterwave baluns are inconveniently long and it is prefer-
able to use either a lumped circuit type or a capacitance-loaded line. The latter,
though physically attractive, has a reduced bandwidth compared with the full
quarter-wave balun. Between 150 MHz and 1 GHz the folded balun or Pawsey
stub is generally the most suitable, (Fig. 17.4). At the highest frequencies the series
inductance of the line AB may be sufficient to unbalance the antenna and the
slotted balun is to be preferred. It can be seen that the balun provides a reactance
shunting the dipole impedance ;Shnitkin and Levy78 demonstrate means of improv-
ing impedance bandwidth by suitable choice of balun and dipole.

xf zOte z in z o ,e x f zA

Fig. 17.2 Sleeve dipole and circuit diagram

Dipoles within the frequency range 30-1000 MHz will normally be self support-
ing and the length/diameter ratio will be smaller than the range plotted by
Schelkunoff for the cylindrical antenna. Resource has then to be made to measure-
ments on 'fat' dipoles such as those described by Brown and Woodward.10 Even
fatter dipoles take the form of cages such as Fig. 17.5 taken from Brown and
Stanier.9 These are 'full-wave' dipoles and it will be noted that at midband the
antenna is only 0.62 of a wavelength. Considerable experimental work on broad-
band dipoles has been done in Germany, for example by Lamberts and Pungs.58
Another variation of the broadband dipole is the bicone (Fig. 17.6). Practical
antennas have cone angles of between 30° and 70°, the larger angles corresponding
to lower characteristic impedances. It should be said here that the main limitation
on the use of such antennas is the band width of the balun used with them. There
are in fact very few applications for such wideband balanced dipoles apart from
antenna test ranges and EMC measurement facilities. Papas and King70 gives
impedance data for conical monopoles.
The folded dipole (Fig. 17.7a) is rarely used as a single element but is commonly
used in stacks for UHF land-mobile base stations, and as the driven element in Yagi
antennas. Apart from its electrical virtues in providing a wide range of input
impedances, the folded dipole has a low DC impedance to ground, which provides
lightning protection and permits the addition of electrical de-icing elements.
For dipoles close to A/2 the impedance Z = (1 4- a)2Zr, where a is a function of
VHF and UHF antennas 729
the radii px and p 2 of the driven and short circuited elements and the spacing d
between centres. When d is large compared with the radii, a is approximately
(Guertler30)

a = lOg(£// )
P2

Z r is strictly the impedance of the equivalent circular dipole, but near resonance the

0=0.0266 L
d=0.375D
—D

• 30i-

,0.266
o-

o
o

-40-

0.222
-80 100
50
resistance,!!

Fig. 17.3 Impedance of sleeve aerials referred to top of sleeve


Measured points in terms of L/X
-+-1 = 0.0266 L - 0 - / = 0.32 L
_ X _ / = 0.214 L -&-1 = 0.428 L
730 VHF and UHF antennas

resistance will not differ significantly from that of the driven element and a first
approximation can be obtained by assuming Zr to refer to the driven element.
When the radiators are of equal radius, the radius R of the equivalent circular radi-
ator is approximately R = yfpd. For unequal radii, the equivalent radius is

l/ f/
R = pv J rjv .

where
P1/P2 = r?
When the length of the dipole is not near X/2, a more generalised approach must be

ZA'B -Z 0i tan9 Zo

Fig. 17.4 Folded balun or Pawsey stub and equivalent circuit

100 110 120 130 1A0 150


f,MHz
571/2in(14 60mm)

9 rods
V4in(6mm)dia
on 8V2in (216mm)
circle

Fig. 17.5 A cage dipole and its impedance characteristic


VHF and UHF antennas 731

adopted. The equivalent circuit shown in Fig. 17.1b shows the basic dipole
impedance stepped up by the ratio (1 + a)2 and shunted by an impedance 2Zf,
where Zf is the input impedance of the short-circuited line formed by each half of
the dipole.

Fig. 17.6 Bicone

Zf = jZ0 tanJC/, where / is the length of the dipole arm, K = 2n/\ and Z o is
the characteristic impedance of the half dipole as a two-wire line. The resulting
impedance for the folded dipole is
2(1 +afZrZf
{l-¥afZr + 2Zf
Two points should be noted about the form of the impedance curve, of which Fig.
17.8 is typical.
(a) The reactance is positive for / < 0.15X, in contrast to the open dipole.
(b) The very high impedance for short element lengths makes the folded dipole of
little practical importance except for near-halfwavelength dimensions or very
narrow-band applications.

Axially-fed dipoles: When the antenna has to be mounted on a support struc-


ture axially-fed versions of the dipole are often more convenient than the balanced
centre-fed form. Three main types will be considered:
(a) Coaxial dipole
732 VHF and UHF antennas

(b) Coaxial dipole with tuned decoupling


(c) Discone

Coaxial dipole: Fig. 17.9 shows the basic coaxial dipole. The radiator A is con-
nected to the inner conductor of the coaxial transmission line, the outer of which is

!"•
o
o

1 +a:1

F ig. 17.7 Folded dipole (a) and equivalen t circuit [b)

2000

1500 1
1000

c| 500

/
0

| / x

1//
-500

-1000 -
0.1 0.2 0.3 0.A 0.5
h/A
Fig. 17.8 Impedance of typical folded monopole

surrounded by a concentric metallic sleeve conductively connected at its upper end.


When the sleeve is X/4 long, it acts as a choke and presents a high impedance
between its lower end and the outer of the transmission line, thus minimising
VHF and UHF antennas 733

current flow on the latter. It is impracticable to obtain a very high Z o for the outer
coaxial line thus formed, so the bandwidth of the choke is limited to the order of
10%. The feed-point impedance of this form of dipole will approximate to that of
the centre-fed dipole of the same diameter, but will be modified by fringing
capacitance at the base of radiator A, especially if a dielectric weather seal is added.

X/A

A/A

Fig. 17.9 Coaxial dipole

Coaxial dipole with tuned decoupling: For frequencies below about 60 MHz, the
outer sleeve may be inconveniently long. Since the aim is to provide a high
impedance at BB, one solution is to shunt a shortened sleeve with a capacitor, the
combination being tuned to resonance at each frequency.
An alternative arrangement is shown in Fig. 17.10. This form is used on military
vehicles in the tactical VHF band 30-76 MHz, where the height of a full dipole
would be impractical but where it is desired to reduce the effect of the vehicle by
minimising current flow on its metal surface. A special triaxial cable is used, coiled
to reduce its height and capacitively tuned between the inner and outer braids
(Brueckmann11)

Discone: Broadband decoupling of the transmission line from the lower half
of the dipole is achieved by making the latter conical. The height of the assembly is
reduced by converting the upper half to a flat disc. The resulting antenna (Fig.
17.11) has an impedance dependent on cone angle and a substantially constant
VSWR above a low-frequency cut-off frequency determined by cone length. Fig.
17.12 shows variation of VSWR with frequency for several cone angles (Nail68).

Coaxial bicone: The broad-band decoupling described above has also been
734 VHF and UHF antennas

applied to an antenna widely used for the military UHF communications band on
ships and at military airfields. The lower section begins as a cone at the feed point,
but is completed by a cylinder to reduce its width. The upper section is usually
conical from the feed point, surmounted by an inverted cone. Both solid and cage
forms exist. A useful reference to such an antenna is Kitchen49 (see also Fig. 17.81).

• skin of
vehicle

Fig. 17.10 Loaded coaxial dipole

Vertically polarised monopoles: Since many forms of vehicle have a metallic


surface on which it is convenient to mount antennas to reduce their projecting
length, the monopole is an important class of antenna for vehicular use. The con-
straints of the vehicle have led to a search for 'low profile' or reduced-height
antennas whose dipole equivalents have no practical application.
When mounted on horizontal circular ground planes they all have essentially the
same radiation patterns as the basic monopole. The influence of ground-plane size
and shape is discussed in Section 17.4.2. Including the A/4 thin monopole or whip
antenna, 16 variants are shown below with brief notes on their significant features.
Dimensions in millimetres are given for a midband frequency of 300 MHz.
(i) A/4 monopole (Fig. 17.13)
Minimum cost, flexible version shown.
Base impedance about 35 £2. LowZ 0 A/4 matching stub at base improves impedance
bandwidth.
(ii) A/4 sleeve monopole (Fig. 17.14)
Impedance at top of sleeve chosen to match transmission line. Sleeve internal
dimensions chosen to give improved impedance bandwidth.
VHF and UHF antennas 735

Improved base strength.


(iii) X/4 broadband sleeve monopole (Fig. 17.15)
Used with streamlined cross-section for low-drag, high-strength aircraft applications.
Capable of 2:1 frequency bandwidth for 2:1 VSWR to 50£2.

Fig. 17.11 Discone

25° discone

400 600 800 1000 1200


frequency, MHz

Fig. 17.12 VSWR curves for discones of various cone angles 0

(iv) Reduced-height top-loaded sleeve monopole (Fig. 17.16)


Strength and aerodynamic qualities as (iii) but reduced height and drag. Bandwidth
similar to (ii). Also used with half top (bent sleeve monopole, see Burberry12).
736 VHF and UHF antennas

(v) Josephson's\/S monopole (Fig. 17.17)


Sleeved and folded monopole with bandwidth comparable to (ii) or (iv).
Sleeved portion can incorporate matching (Josephson42).

250

Fig. 17.13 \l'4 monopole

Fig. 17.14 Sleeve monopole

250

Fig. 17.15 Broadband sleeve

(vi) Printed X/8 monopole (Fig. 17.18)


Double-sided PCB gives element suitable for encapsulation in highly streamlined
form. Element dimensions similar to Fig. 17.17.
VHF and UHF antennas 737

(vii) 'Low profile' or transmission-line antenna (Fig. 17.19)


Capacitance loaded, shunt feed, bent monopole.
Instantaneous bandwidth about 1%, but tunable over octave with some loss of gain
at low-frequency end of band. For narrow bands may be made self resonant, length
250 mm at 300 MHz. Gain then comparable to (i). Pattern circularity degrades with
increased capacitance.

125 125

Fig. 17.16 Top-loaded sleeve monopole


Dimensions in mm for 300 MHz

100

125

Fig. 17.17 Josephson's \/8 monopole

(viii) Annular slot (pill box) (Fig. 17.20)


50% bandwidth possible with dimensions as shown. Can be made much smaller with
corresponding loss of bandwidth, e.g. 66 mm square, 1.77 mm deep, 2\% bandwidth.
(ix) Annular slot (planar) (Fig. 17.21)
Dimensions similar to (viii) but antenna is flush mounted instead of projecting.
738 VHF and UHF antennas

(x) Cone (Fig. 17.22)


Broadband with defined low-frequency cut-off.
Optimum cone angle 30°-50°.

r 11
1 1— 1 1
1 1 1 1

1i tJ
1 1 .1
125

i 1 1

\c J 7
/ / / / / / / / / / / / / /
Fig. 17.18 Printed X/8 monopole

L - 150

Fig. 17.19 Low-profile or transmission-line antenna

500 dia.

Fig. 17.20 Annular slot (pillbox): cross-section

(xi) Grounded discone (Fig. 17.23)


Cross-section elliptical. Used as fin cap antenna in aircraft.
Resonant frequency lower than given by cone dimensions. May be regarded as
variant of (iv) with enhanced bandwidth.
(xii) Aerodiscone (Fig. 17.24)
Similar to (xi) but with earthing pins for matching (see Barbano6)
VHF and UHF antennas 739
(xiii) Top-loaded folded monopole (Fig. 17.25)
Used at 150 MHz for railway vehicles. Provides moderate bandwidth in short, robust
antenna. DC grounded.

500 dia.

Fig. 17.21 Annular slot {planar)


(a) cross-section
(b) enlarged view

(xiv) Helical monopole (Fig. 17.26)


Usually enclosed in flexible dielectric sleeve and much used on pocket radios. Short,
fat helix gives better bandwidth than long, thin one which approximates to (i).
(xv) Hula hoop (Fig. 17.27)
A variant of (vii) having more circular radiation patterns less dependent on amount
of end loading. Bandwidth properties as (vii).
Loop circumference less than 200 mm at 300 MHz. DC grounded.
(xvi) Notched plate (Fig. 17.28)
Flat plate with self-resonant notch in one vertical edge. Notch may be folded as
shown and tapered for broad band. Essentially omnidirectional. DC grounded.
Three of these forms are described below in more detail.

(a) 'Low profile' or transmission-line antenna (Fig. 17.19): Wanselow and


Milligan89 give a good description of this antenna and a number of variants which
may be tuned over wide frequency bands. The antenna may be considered as a
740 VHF and UHF antennas

short vertical radiator with horizontal top loading. In its simplest form for single-
frequency operation no capacitor is necessary. In this configuration the gain can be
comparable to the normal A/4 monopole. Experience shows that for maximum
efficiency large-diameter elements are necessary and good grounding of the vertical
element is of prime importance because of the very low radiation resistance of the
basic radiator. This will become apparent from the following analysis.

ill

Fig. 17.22 Cone

Fig. 17.23 Grounded discone

Referring to Fig. 17.19, quarter-wave resonance will be achieved when


0! + <92 = 90°
VHF and UHF antennas 741
where dx is the electrical length between the drive point and the capacitor and 62 is
the length of open-circuited transmission line to which the capacitor is equivalent.
Now 7 v

= cot where X
2irfC
and
Z o = 138 log 10

matching pins

Fig. 17.24 Aerodiscone

70

F ig. 17.25 Top -loaded folded m on op ole


Top may be circular, elliptical or rectangular

For transmission lines of this form the phase velocity can be considered equal to
the speed of light, so that the physical length I is given by
Bx - 360I/X
The vertical radiator, when it is short compared with the wavelength, can be con-
sidered as a constant-current element of radiation resistanceRo ~ I515(h/X)2 (see
Schelkunoff and Friis77 p. 309), where h is the height of the radiator. This is the radi-
ation resistance for zero end capacitance; as the capacitance is increased the current
742 VHF and UHF antennas

through it increases and the radiation resistance is reduced to Rr = R0[l — (/ c // 0 ) 2 ]

T V
-* = cos (90° -62) = sin cot"1 - *

140

Fig. 17.26 Helical monopole

Fig. 17.27 Hula hoop


Circumference < X/4
Height « 0.05X

It should be noted that, as C is increased, the antenna changes from an essentially


vertically polarised radiator to a half-loop of near constant current and the elevation
radiation pattern changes accordingly.
Although the instantaneous bandwidth of the low-profile antenna is small, of the
order of 1% for acceptable VSWR limits, it can be designed for tuning over an
octave bandwidth with no other adjustment than the capacitance.
Unlike the dipole and monopole which have higher-order resonances, the
VHF and UHF antennas 743
impedance of the low-profile antenna displays a single resonance and the antenna
has inherent band pass characteristics.

(b) Annular slot: Two forms of annular slot are used (Figs. 17.20 and 17.21).
They differ in the position of the slot, and hence in application. Whilst the flush
mounted type (Fig. 17.21) may be more familiar, there are many applications for
which the pill box (Fig. 17.20) is more suited since it does not require a large cut-
out in its mounting surface.

250

\\vv\u
Fig. 17.28 Notched plate
f V\\

This discussion applies only to small annular slots having diameters up to 0.6X.
Within this range the radiation patterns are virtually those of a vertical monopole.
Johnson and Rothe39 have considered this antenna as the slot analogue of a
circular loop, the slot being backed by a radial transmission line short-circuited at
its centre. The slot is loop coupled into a coaxial line. Using the nomenclature of
Fig. 17.21, the slot impedance is seen as a resistance Rs shunted by capacitance C,
the latter given approximately by
C = 0.17816D loge Djh picofarads, D and h in centimetres
Rs is derived from the equivalent loop resistance
(1207T)2
Re by the equation Rs =
2/ve
Fig. 17.29 gives values of Rs against slot diameter D
The slot width has little effect on Rs
The impedance Z ( r ) of the short-circuited radial transmission line is in parallel
with the slot capacitance and is given by
744 VHF and UHF antennas

= _ h J0(Kr)N0(Kr0) - J0(Kr0)N0(Kr)
(r) }
r Jx(Kr)N0(Kr0)-J0(Kr0)Nx(Kr)
where / 0 , Jx and A^o, Nx are Bessel functions of the first and second kind respect-
ively, of orders 0 and 1. Z ( r ) is chosen to resonate with the slot capacitance. At
resonance the equivalent series resistance, R's of the slot becomes

1000

500

100

50

10
'0 0.5 D 10 1.5
X
Fig. 17.29 Radiation resistance of an annular slot radiating to one side only

Considering now the loop formed by the inner radius of the cavity, the feed line
and the top and bottom surfaces of the cavity, the mutual impedance of this loop is
given by
N0(Kr0)J0(Kl)-J0(Kr0)N0(Kl)
M =
N0(Kr0) Jx (Kr) - J0(Kr0)Nx (Kr)
where JU is the permeability, normally unity. The input impedance R at the feed
point is given by RR'S = (coM)2 hence
VHF and UHF antennas 745

N0(Kr0)J0(Kl) -J0(Kr0)N0(Kl)
R — R
S
[N0(Kr0) J0(Kr) -
This neglects the inductance of the feed wire for which capacitance compensation
may be required.
Cumming and Cormier22 give some useful design curves.
Whilst the design data given above refers to circular antennas, a square slot has
similar radiation characteristics and may be easier to construct.

(c) Notch antenna: The short, capacitance-loaded notch has been described in
Chapter 16. For frequencies above about 100 MHz, the use of self-resonant notches
provides wide bandwidth in convenient mechanical form. Fig. 17.28 shows a notch
on a printed-circuit board, the notch being coiled up to avoid a plate width greater
than X/4 which would introduce directivity in the azimuth plane. Experience shows
that the coupling between adjacent turns is negligible provided that the spacing is of
the order of four times the slot width.

feed point

L2

mouth

Fig. 17.30 General notch and equivalent circuit

Two alternative methods of considering the general notch antenna, (Fig. 17.30)
will be discussed. Johnson40 considered it as the magnetic analogue of a monopole
on the edge of a metal sheet. Using Babinet's principle, the mouth or open-end
746 VHF and UHF antennas

resistance of the equivalent notch becomes


240TT
3[C(Kl) - cos (Kl)S(Kl)]2
where C and S are Fresnel Integrals.

admittance
.at open end of notch

shunt
capacitance -§
required

Fig. 17.31 Circle diagram of notch aerial admittances.


f j . . . f4 typical feed-point admittances

This resistance is shunted by a reactance due to the short-circuited slot


transmission line which is the notch.
An alternative way of considering the notch is as a slot in a flat sheet folded
about its centre line. It is clear that, whichever approach is adopted, the open-end
impedance of a quarter-wave notch will be of the order of 400 O. A matching point
to a coaxial line can be found in the same way that a slot can be tapped down to an
appropriate impedance.
Fig. 17.31 shows the process graphically on a Smith chart, assuming the slot
transmission line to have Z o = 100. Broadband operation can be obtained by using
a tapered notch, narrowing from the open end, (Fig. 17.32). In a three-dimensional
version, for example in an aerofoil section, a physically parallel notch will have a
naturally tapered Z o .
VHF and UHF antennas 747

Horizontal loops: A horizontal loop has a substantial omnidirectional pattern


provided its diameter is less than about A/3. For the frequency ranges of interest, it
is usual to use single turn loops only; the radiation resistance of such a loop is

R ~ 31,000 -j where S = area of loop


A
The inductance of a single-turn loop is
b
L = ixb log —
a
where a is the radius of the wire and b is the radius of the loop. Use of such loops
is somewhat limited, although, as will be shown subsequently, they are used in
arrays.

Fig. 17.32 Two-dimensional wideband notch


T— Sheet thickness
= 7.5

Several loops may be connected in parallel in the horizontal plane (Fig. 17.33),
giving an arrangement with substantially in-phase currents and good omnidirectional
coverage. Normally four loops are used, but three are possible at some sacrifice of
circularity.
One variation on the loop is the short notch cut into a horizontal disc (Fig.
17.34). The same constraints on diameter apply as for the loop. Since the notch is
short, it requires capacitive loading at its mouth and has a narrow (1%) instan-
taneous bandwidth although capable of tuning over at least 10% band. If the disc is
748 VHF and UHF antennas

made as a discus a structurally attractive antenna results which can be mounted on


a centre metal pole.

Turnstile or crossed dipole: A pair of orthogonal centre-fed dipoles will give a


good approximation to an omnidirectional pattern if there is a 90° phase shift

feed point

Fig. 17.33 Alford loop

< A/3

F ig. 17.34 No tched disc

between their input currents whose amplitudes are equal. Whilst it would be possible
to feed from a common transmission line with an additional X/4 to one dipole this
leads to unequal power sharing since the junction impedances are unequal, unless
these impedances are conjugate. Such an arrangement results in a narrow bandwidth,
and a better system results from connecting the dipoles by equal transmission lines
VHF and UHF antennas 749
to a 3 dB coupler. In this form of the antenna, the pattern circularity is determined
by the VSWR at the coupler input ports.

Slotted cylinder: This antenna consists of an axial slot cut in a hollow conduc-
tive cylinder, the slot being fed from a parallel or coaxial line. The antenna may be
compared with a vertical stack of close-spaced horizontal loops connected to a
2-wire transmission line short-circuited at top and bottom.

A/A

to phasing
network

Fig. 17.35 Stack of \/4 elements on mast

17.2.2 Stacked elements

Vertically polarised: The problem with a stack of vertical dipoles is to dispose


the cables to minimise their effect on the radiation pattern. One method is to
750 VHF and UHF antennas

support each dipole on a horizontal yard arm projecting from a vertical mast, the
spacing being chosen to give the least perturbation. Carter16 gives methods of calcu-
lating the resultant field. This may result in an inconveniently large structure, and
one alternative is to use fat dipoles wrapped around the support mast. Fig. 17.35
shows one such arrangement using X/4 elements. It is only necessary to feed the end
pairs. Three pairs of elements, outer pairs fed, give 4- 3 dB gain on a A/2 dipole,
whilst 14 pairs, with only the outer pairs fed, give + 6 dB gain.

Fig. 17.36 Stack of X/2 elements on mast

An alternative arrangement using full-wave dipoles is shown in Fig. 17.36. As


might be expected, the dipole elements have intrinsically wide bandwidth, and with
proper adjustment of the choke dimensions an octave bandwidth can be achieved
for each element. The bandwidth for a stack will, of course, depend on the number
of elements, the spacing, and the feed system. In this respect it should be noted
VHF and UHF antennas 751

that a centre-fed stack with vertical symmetry will give maximum horizontal plane
radiation, whilst a progressively fed system can give a beam tilt which will be fre-
quency-dependent.

Fig. 17.37 Stack of dipoles round mast with progressive phasing

A vertical dipole close to a vertical mast will have a broad unidirectional pattern.
Two or three elements spaced round the mast will create an approximately omni-
directional pattern. To reduce their mutual coupling the elements may be staggered
in the vertical plane as in Fig. 17.37. If series, in-phase, feeding is used, the dipoles
will normally be folded to give a satisfactory match to a 50 £2 line.

Horizontally polarized: Stacking of horizontal elements on a vertical mast is


comparatively simple since the antenna and the mast are orthogonal. The turnstile,
the loop and the slotted cylinder can all be used, the last sometimes being used for
VHF TV transmitting antennas because the cylinder dimensions are compatible
with mast size for many applications.
752 VHF and UHF antennas

A modification of the turnstile known as the super-turnstile makes use of


elements known as batwings (Fig. 17.38). Each element may be considered as a
stacked pair of horizontal dipoles or, alternatively, as a slot in a shaped conductive
sheet. The shaping of the sheet improves the impedance characteristic of the
element.

A/2

F ig. 17.38 Basic ba twing elemen t


Shaded area may be mesh or grid of rods

Fig. 17.39 Vertical dipole and cylinder

Transmitting antennas for television: Extensive arrays of dipoles placed around


support masts are used for television broadcasting. The arrays may be of either
polarisation. The feeds to the dipoles for main-station transmitting aerials are pro-
gressively phased in order to tilt the beam down towards the horizon, and pertur-
bations are introduced into the dipole feeds to fill the zeros in the vertical radiation
VHF and UHF antennas 753

pattern and so avoid gaps in the service area (Platts71). Main station antennas are
nearly always omnidirectional,but smaller antennas used for low-power relay stations
often have directional horizontal radiation patterns to satisfy particular service
requirements (Millard65).
0.24 A

dipole

Fig. 17.40 Horizontal dipole and cylinder

17.2,3 Effects of mounting structure


The simplest form of mounting structure is the mast in the form of a vertical
cylinder. Carter16 shows how to calculate horizontal radiation patterns for
vertical and horizontal Hertzian doublets. Typical patterns are shown in Fig. 17.39
and Fig. 17.40.
This method may also be used to compute the effects of masts of polygonal
cross-section provided that the width of each face is small compared with the
wavelength. The use of equivalent diameter cylinders is described by Knight,51 who
also discusses other methods for appropriate sized masts. Thus the infinite plane
method may be used for polygonal masts having faces wider than 2X, but fails to
account for the edge diffraction on smaller masts. For these, either the induced-
current method or formulas involving wedge diffraction can be used. Fig. 17.41
compares patterns calculated by these four methods with measured patterns for
dipoles A/4 from the face of a triangular mast of side IX.
Patterns for elements on or adjacent to elliptical cylinders are considered by
Sinclair.83

17.3 Directional antennas

Directivity can be achieved by the combination of separate elements in appropriate


phase and amplitude relationships or by the method of reflection from conducting
surfaces. The latter class, using what are known as aperiodic reflectors, may use
only a single driven element and are therefore among the simplest systems to
construct.
754 VHF and UHF antennas

17.3,1 Aperiodic reflectors

Single sheet: A centre-fed dipole parallel to an infinite conducting surface at a


distance S from it can be replaced by a pair of parallel dipoles spaced 25, the second

30° 60° 90° 120° 150° 180

0.8

V
4 I

0.6

0.4
*Cv

0.2 -

0
30° 60° 90° 120° 150° 180°

Fig. 17.41a Theoretical and measured horizontal radiation patterns (HRPs) of a vertical
dipole spaced 0.25\ from a triangular mast 1.0\ wide.
HRP calculated by Carter's method
HRP calculated by the infinite-plane method
— - — HRP calculated by the induced-current method
HRP calculated by the diffraction method
o o o o Measured points (averaged over 360°)

Table 17.1 Gain of dipole and finite, square, flat sheet reflector
Square of side L/A Spacing of A/2 dipole, S/X Gain w.r.t. dipole, dB
1 0.2 5.2
Large 0.215 5.8
1.5 0.775 7.0
or image dipole having a phase of 180° referred to the first dipole. The radiation
patterns, impedance and gain can be computed for the dipole pair, taking into
account the mutual impedance between them. If the conducting surface is finite,
diffraction from the edges of the sheet must also be taken into account. The edges
parallel to the dipole have the greater effect, that of the perpendicular edges being
VHF and UHF antennas 755

negligible if their spacing is slightly greater than the length of the dipole. Fig.
17.42 shows variation of radiation pattern with sheet width and indicates that a
comparatively small sheet can be an effective reflector. Fig. 17.43 shows variation
of impedance and gain with spacing for an infinite sheet and Table 17.1 gives some
typical gain figures for finite sheets. It will be noted that, if forward gain is the only
criterion, there is little advantage in increasing the sheet size beyond a certain

30° 60° 90° 120 150° 180°

0.8

-* 0.6
LU

X\
\
0.A

0.2

0
30° 60° 90° 120° 150° 180°
0
Fig. 17.41b Theoretical and measured horizontal radiation patterns (HRPs) of a horizontal
dipole spaced 0.25\ from a triangular mast 1.0\ wide
HRP calculated by Carter's method
HRP calculated by the infinite-plane method
HRP calculated by the induced-current method
HRP calculated by the diffraction method
o o o o Measured points (averaged over 360°)

minimum dependent on spacing. The front-to-back ratio for a dipole at X/4 in front
of a sheet of width 2x is given in Fig. 17.44 and here, too, there is little sense in
using sheets wider than 2X.

Corner antennas: The gain of the dipole and single sheet is clearly limited to the
order of 6 dB. Higher gains can be achieved by folding the sheet to form a corner. A
756 VHF and UHF antennas

90° corner can have a gain of up to 13 dB whilst with a 60° angle a gain of 14.5 dB
is possible. As Fig. 17.45 indicates, these gain figures require a larger apex-antenna
spacing than is normally acceptable physically. As the corner angle is decreased the
length of the side has to be increased to give a good approximation to the
theoretical performance of infinite-length sides. For any reasonable bandwidth a
90° corner with IX aperture and spacing 0.35X represents the best engineering
solution. It may well be physically more desirable to stack two 90° corners one
above the other than to build a 60° corner with 1.25X spacing to achieve the same
gain.

axis of
symmetry

0.25 A

D/X
-2.215

field strength
axis of
Hsymmetry

Fig. 17.42 Effect of sheet width on pattern of dipole and finite sheet

It is not necessary to construct a screen or corner of solid metal, a mesh or grid


of wires being adequate provided that one set of wires is parallel to the electric
vector. If the wires of radius a are spaced at a distance d, the transmission coefficient
for an infinite screen is
VHF and UHF antennas 757

7 = (2dl\)\oge(dl2ua)
Front-to-back ratio is much more sensitive than is forward gain to the wire spacing
and diameter. In some instances some backward radiation is desirable; a striking
example is the 24-dipole corner array used as the localiser antenna in high-accuracy

0.1 0.2 0.3 0A 0.5 0.6 0.7 0.8 0.9 1.0


R/A
a
> 9-
6 '"o A

ir2
0.2 0.4 0.6 0.8 1.0
P/A

Fig. 17.43 Impedance and gain of dipole parallel to infinite sheet

50

A0

CD 30

20
2 3
2x /A
Fig. 17.44 Front/back ratio for dipole and finite sheet

VHF instrument landing systems where backward radiation provides outward


guidance in case an aircraft has to overshoot the runway. Operating in the 108-
112MHz band, this has a screen of 1.5 mm-diameter wires spaced 150 mm, giving
a back beam approximately 15 dB below the forward beam. This is rather more
shielding than a plane grid would provide since the corner sides are oblique to the
rearward direction.
While 'fish-spine' reflectors as Fig. 17.46 are convenient for single-point sup-
ported corners, wooden or metal frames covered with 'chicken-netting' are equally
758 VHF and UHF antennas

effective as reflectors. It should be noted that at the shorter wavelengths a stiff


frame with mesh covering may well be lighter, cheaper and as robust as a 'fish-spine'
construction.

0.1 0.2 0.3 0.4 0.5


dipole-apex spacing A

Fig. 17.45 Gain of corner antennas

Short backfire antenna: Fig. 17.47 shows the essential features of the short
backfire antenna which was first described by H. W. Ehrenspeck.26 A good
theoretical explanation has not been forthcoming, but it seems probable that the
mechanism is as follows:
Consider initially a plane sheet and dipole feed. This belongs to the single-sheet
class of antenna discussed above and its gain cannot significantly exceed 6 dB. It
can be considered as a dipole pair plus the vector addition of radiation from the
sheet. A similar analysis would apply if the feed were a dipole-dish combination
but the illumination of the sheet would then be altered.
The addition of a rim has two effects. It alters the radiation due to the sheet by
modifying the edge condition and superimposes another radiation component at a
different point in space. It is not clear whether the edge or the surfaces of the rim
are the more significant; if the former, a loop similarly placed should have a similar
effect. There is no evidence from the literature that this has been tried.
The large increases in gain obtained by the addition of rims are only possible
provided that there is adequate illumination of the reflector; so this type of antenna
is limited to reflectors of the order of 2-2.5X diameter maximum. Narrow-band
gains of the order of 18 dBi are reported, but since these require optimisation of the
rim at specific frequencies, the broadband gain is probably of the order of 12dB.
Larger reflectors, with appropriate increase in gain, can be used with multiple feeds.
For circular polarisation crossed dipole- or tripole- and disc combinations may
be used.
VHF and UHF antennas 759

Typically the dimensions in Fig. 17.47 might be d = 2A, hx = h2 = 0.25X, 2r =


A/2, W up to 0.6A.
Nothing has been recorded about the cross polarisation of such antennas. Side-
lobe levels are typically more than 20 dB down and could probably be improved at
the expense of gain for which most systems are optimised.

> 0.6 A

metal or dielectric
support

Fig. 17.46 Corner antenna with fishspine reflector

dipole

Fig. 17.47 Short backfire antenna

A recent paper suggests that the gain and sidelobe level can be improved by the
use of rim corrugations Fig. 17.48. Fig. 17.49 shows measured E-plane patterns.

Shaped reflectors: The use of singly- and doubly-curved reflectors in this


760 VHF and UHF antennas

frequency range is very much limited by physical constraints. At 300MHz, for


example, a lOft-diameter paraboloid with 55% efficiency will provide a gain of
17 dB. A more effective structure for the same gain can be achieved in other ways,
such as an array of Yagis, so that there should be some very real advantage before a
curved reflector is used. Except in special circumstances the gains required in this
frequency band are not sufficiently great for the main advantages of curved
reflectors to be dominant. This is particularly true of pencil-beam antennas; the use
of long parabolic cylinders for fan beams is common for surveillance radars and for
radar astronomy.

D
3A A/2
D

JUI
A/4i A/4

Fig. 17.48 Short backfire with improved sidelobe

For most purposes dipole/reflector combinations are used as feeds. A dipole


alone is inefficient since half its energy is radiated forward and only half its energy
directed at the dish. The reflector is usually 0.25 X in front of the dipole and it may
be a disc of about 0.5 X diameter or a rod or strip about 0.5 X long. Linear feeds do
not have equal E- and H-plane radiation patterns and the resulting dish patterns
must therefore be unequal. In this respect, a focus outside the plane of the reflector
edge will be an advantage since the feed cannot provide wide-angle illumination.
The factors to be considered in the design of a small paraboloid antenna are:
(a) The relative importance of maximum gain for a given diameter and minimum
sidelobe level
(b) Bandwidth
(c) Limiting VSWR
The phase centre of the feed should be at the focus of the dish, but there is some
VHF and UHF antennas 761

uncertainty with dipole-disc and dipole-reflector feeds regarding the position of


the phase centre. Reason suggests that it must lie between the two elements and
some axial adjustment is usually necessary. With dishes whose diameter D is small in
terms of wavelength, i.e. D/X < 2, and hence with small focal length F/X, the gain
and sidelobe level change rather rapidly with frequency, and wide bandwidth is not
possible with a single feed position.

60 50 40 30 20 10 0 10 20 30 40 50 60

Fig. 17.49 E plane patterns


With corrugations
Without

The illumination of the dish has to be tapered to reduce sidelobes and loss of
gain due to spillover. Typically an edge illumination lOdB below peak would be
used if maximum gain is the main criterion, but — 20 dB would be preferred for
minimum sidelobe levels. The distance F from the focus to the dish increases with
angle off-axis, so that energy at an angle 6 is attenuated by 20 log (sec2 0/2) dB,
compared with the energy on axis. This has to be considered in the design of the
feed radiation pattern, and it will be found that, for small F/D, it is difficult to
obtain an appropriate feed pattern with a dipole/reflector combination. In conse-
quence, the gain will rarely be more than 65% of that obtainable from a uniformly
illuminated dish of the same size.
Energy from the feed is reflected back into the feed from the area adjacent to
the vertex of the dish. The reflection coefficient due to the dish is F = gXjAnF,
where g is the axial gain of the feed. ^ _ _
A small circular plate radius yFX/3 placed A/24 in front of the vertex will
reduce the mismatch, but is rarely used now because of its narrow bandwidth.
The slotted feed (Fig. 17.50) which is often used for dipole/reflector combi-
nations provides an impedance transformation whose effect has not always been
762 VHF and UHF antennas

recognised. If the inner of the coaxial line is continued into the slotted section, it
is obvious that its effect is modified by the short circuit to one side of the slotted
section. In effect the dipole impedance is transformed by a line of impedance 2Z 0
shunted by a short circuited section also of 2Z 0 as shown in Fig. 17.51a and b. An
assumption has been made here that the slot is X/2 long and itself has no effect on
the dipole impedance. This is true so long as the slot is not covered by a dielectric
which would modify its effective length.

^reflector

slot both sides


of cylinder

Fig. 17.50 Slotted dipole feed

17.3.2 Parasitic elements


Any radiating element not driven by a feed line but coupled to a directly fed
radiator can be regarded as a parasitic element. Although dipole-type elements are
the most widely used, it is also possible to have parasitic loops, slots and notches.
Some of these occur in common mechanical structures — a motor vehicle, for
example, has slots between bonnet and body, and boot and body, and the windows
act as parasitic slots. The moving control surfaces of aircraft form, with the struc-
tures to which they are attached, a series of slots and notches with individual
resonances. Fig. 17.52 shows a variety of parasitic elements on an aircraft.
The simplest combination of elements consists of one driven element and one
parasitic. The radiation pattern, gain and impedance are determined by the follow-
ing factors:
(i) Magnitude of current in the parasite
(ii) Phase of energy re-radiated by parasite
These are themselves controlled by the physical spacing and by the reactance of the
parasite, itself a function of element size and cross-section. Little has been
published on the analysis of slot, notch and loop parasites but the treatment is
VHF and UHF antennas 763
88
similar to that for dipole systems described by Walkinshaw. In this reference radi-
ation pattern, gain and impedance are calculated in terms of parasite spacing and
self reactance. A parasite shorter than X/2 is known as a director, since the maxi-
mum radiation is on the same side of the driven element as the parasite. A parasite
longer than X/2 is referred to as a reflector as the maximum radiation is directed
away from it. Walkinshaw's paper shows that, if a single parasite is used, a director
(negative reactance) is better at spacings of the order of 0.1X whilst for larger
spacings a reflector is to be preferred. Reference to tables of mutual impedance
will show that, whatever the reactance of the parasite, its effect will be negligible
at spacings much in excess of X/2. It should be noted that the spacings and lengths
for maximum forward gain are not the same as those for best front/back ratio.

dipole
e

/ / \
/ N
2Z 0
Zo 2Z 0
02 B

ZA

Ao-

Fig. 17.51 (a) Schematic ofslotted feed; (b) Equivalent circuit


Xx = 2/Z 0 tan dl
X2 - 2/Z 0 tan d2
Zj± — Dipole/reflector impedance

Gains compared with a half-wave dipole of 5.5 dB for a director and 5.2 dB for
a reflector can be achieved. High gain at close spacing results in very low input
resistance to the driven element, of the order of 10-15 0 at 0.1X spacing. A folded
dipole (see Section 17.2) is often used to improve the match to the transmission
line, but small bandwidths are usual with such small spacings. If operation over
bandwidths up to 20% is required, a dipole and reflector combination at 0.15 X to
0.2X is to be preferred.
764 VHF and UHF antennas

The use of full-wave dipoles and tuned reflectors is uncommon, but was the basis
of a wide range of antenna systems developed in Germany for such diverse uses as
500MHz airborne radar systems and 41-68 MHz point-to-point links. The basic
element of the latter system is shown in Fig. 17.53. It consists of a 'fat' (D — 0.07 A)
driven element, 0.7X long with a rod reflector at 0.3X spacing. The use of full-
wave dipoles permits metallic supports at the centre point of each arm, which is
electrically dead. The reflector and dipole spacings were adjusted to give compro-
mise performance over the band 41-68 MHz. A VSWR of 1.3 at 60 £2 was achieved
with a gain of 12 dB over the band.

F ig. 17.52 Parasitic elemen ts on aircra ft


A: Tailplane-elevator notch
B: Elevator-fin notch
C: Fin-rudder notch
D: Wing-aileron notch
E: Aileron-flap notch
F: Flap-flap notch
G: Flap-fuselage notch
H: Pitot static probe
J: Ice detector

Because of the need for a metal surrounding surface it is not possible to use slots
and notches in end-fire as is done with dipoles. They can only be used in broadside
array to skew the beam from the normal broadside position. Such parasitic slot
VHF and UHF antennas 765

systems have been used on aircraft to provide unidirectional radiation, the tuning of
the parasitic slot being achieved by a combination of its dimensions and those of
the backing cavity.
Similarly, whilst it is possible to have an end-fire parasitic array of horizontal
loops, there is no advantage in using parasitic vertical loops — a dipole system
would be smaller and more directional.

JL JL 7 3A

0.625A

3 reflectors
omitted for clarity

0.7A

Fig. 17.53 Stack of 4 full-wave dipoles and reflectors


Three reflectors omitted for clarity

Yagi-Uda array. The designs of Prof. Uda of Japan were first described in
English by Mr. Yagi, so this type of parasitic array should bear their joint names. It
is a logical extension of the use of parasitic dipoles with one driven element, in end-
fire array.
As Walkinshaw88 demonstrates, the most effective arrays use a number of
directors but only a single reflector. Whilst in theory the forward gain can be
increased indefinitely by increasing the number of directors, in practice there is
little to be gained by using more than 12. If higher gains are required it is better to
use arrays of Yagi-Uda arrays; this has the double advantages of:
(a) a structurally better system
(b) independent control of sidelobe levels by adjustment of array spacing. The
diminishing returns from multi-director Yagi-Udas is to be expected since successive
elements have smaller amplitude currents induced in them.
In general, the addition of parasitic elements reduces the radiation resistance of
the driven element so that a folded dipole is necessary to match the impedance to
normal transmission lines. There is one exception in the combination of one driven
766 VHF and UHF antennas

element and four directors each spaced 0.2X where the input impedance is approxi-
mately 45 O and no folding is required.
For mechanical convenience the parasitic elements are normally mounted on a
central metal tube which, since it lies along the neutral axis of the system, does not
perturb the performance apart from a slight modification of the effective length of
each element.
Normally Yagi-Uda antennas have narrow bandwidths, of the order of 2% for
optimum gain arrangements. Some considerable improvement in bandwidth can be
obtained at the sacrifice of some gain. One method of doing this is to design the
directors for the upper end of the desired band and the reflector for the lower end
of the band. Two five-element Yagis (one reflector, three directors and one driven
element) of comparable size, one optimised and one broad band, had gains respect-
ively of 8 dB over a 9% band and 4 dB over a 60% band, referred to a half-wave
dipole.
Since the effect of a parasitic element is a function of its length and diameter,
any change to these will affect its performance. One natural hazard is the formation
of ice, which modifies the effective diameter of the element and hence its reactance.
Experiments in Germany showed that it was possible to reverse the direction of the
main beam by ice loading, the effect being more severe on the thinner elements of a
500MHz Yagi than on those of a 220 MHz design. It seems unlikely, therefore, that
icing would be a severe problem below the UHF band.
The dielectric constant of ice in the VHF and UHF bands is approximately 3. A
fair simulation of the effect of ice can be achieved by putting glass fibre tubes
having a diameter equal to the anticipated ice, over each element of the array.

Fig. 17.54 Quad antenna derived from pair of dipoles and reflectors

Cubical quad: This antenna is little known outside the amateur ranks but has
some practical advantages over the simple Yagi. The derivation of the antenna from
a stack of two horizontal Yagis is shown in Fig. 17.54. Fig. 17.55a shows a typical
3-element array with sides of approximately X/4. In order to tune the reflector and
director a short-circuited and open-circuited stub, respectively, can be inserted in
the arms parallel to the feed point Fig. 17.55Z?. The gain of the 'quarter-wave' quad
VHF and UHF antennas 767
with three elements is 7-7.5 dB, but a 'half wave' quad (sides approximately A/2)
can have a gain of 12.5 dB.
The quarter-wave quad is attractive in its reduced width compared with a Yagi,
and is said to possess wide bandwidth and to be dimensionally non-critical.

F ig. 17.55 Two forms o f 3-elernen t quad

Fig. 1756 Log periodic dipole antenna


R2 R n
^n+i
T = =
^

17.3.3 Log periodic antennas


The broadband properties of the log periodic antenna arise from its operation
over a number of small bands sufficiently closely spaced to provide essentially
uniform performance over a wide range of frequency. The operation can be most
768 VHF and UHF antennas

readily explained by consideration of the log periodic dipole antenna shown in Fig.
17.56. This is only one of many forms of log periodic antenna but is one of the
more suitable physically in the VHF and UHF bands.
The antenna consists of a number of centre-fed dipole elements connected to a
twin transmission line so that the phase is reversed at each connection relative to
the adjacent elements. The element lengths and spacings are related as shown. The
antenna is fed at its high-frequency end and the twin transmission line is short-
circuited behind the longest element. For convenience, a coaxial cable can be taken
through one of the hollow twin lines.
Consider the frequency for which the «th element is half-wave resonant. This
element will have the lowest impedance of the array and will therefore abstract
more energy than any other element. It will be obvious that the smaller (higher-
frequency) elements will present a higher impedance and the energy abstracted will
progressively increase to the nth element and decrease beyond it. The active
part of the antenna is restricted to a few elements either side of the resonant
one. The possibility of effects due to 3X/2 elements can be excluded since very
little energy travels along the transmission line beyond the half-wave resonant
element. Increasing the frequency merely moves the active section towards the
high-frequency end of the antenna. The antenna can, therefore, be considered as a
series of short antennas optimised at the resonant frequencies of successive
elements. The degradation of performance between optimised frequencies depends
on the periodicity of the antenna and has to be weighed against an excessive
number of elements. The lowest operating frequency will be determined by the
length of the largest element, the highest frequency by the smallest. Energy travel-
ling beyond the longest element will be reflected back by the short circuit and the
antenna will not exhibit the single-lobed radiation pattern which is characteristic of
the operating frequency band. In the range 30-1000 MHz practical arrangements
of elements do not limit the upper frequency performance; a typical arrangement
is shown in Fig. 17.57.
The significant factors are r, the ratio between adjacent element lengths and
spacings, and a, the spacing in wavelengths between one element and its next,
smaller neighbour. Currel15 derives design curves relating r and o to gain and has a
useful series of nomographs giving structure bandwidth, structure length, and num-
ber of elements in terms of these two parameters. The useful range of r is shown to
be between 0.76 and 0.98 for values of o between 0.06 and 0.20. Gains are between
8 and 12dB, the latter only being achieved for values of r near the upper limit,
where clearly the active region includes more elements.
The mean resistance Ro of the antenna is shown to be a function of the
transmission-line characteristic impedance Z o which has little effect on the direc-
tivity. As might be expected there is a reduction in directivity with thinner
elements since their impedance change is more rapid with frequency. The VSWR
referred to Ro tends to rise withZ 0 and with decreasing o. Typically VSWR values
of 1.4 are achievable for a wide range of parameters. The operating bandwidth is
given approximately by
VHF and UHF antennas 769
B = BjBa
where B8 is the structure bandwidth and B^. is the bandwidth of the active region.
£9 = T1-N where N is the number of elements and Bar lies in the range 1.2 to 30
being lowest for large r.
Further design details can be found in Smith.84

F ig. 17.57 A ttachmen t of dip ole elemen ts

The details given above refer to structures with half-wave straight dipoles. Mayes
and Carrel63 have shown that vee-dipoles operating in the 3X/2 and higher modes
can provide higher gains; straight dipoles would have undesirable multilobed pat-
terns and cannot be used. The arms of the vee are sloped towards the high-frequency
end of the structure. Gains of 12 dB and 18 dB, respectively, have been reported for
3X/2 and 5X/2 vee dipoles but the increased disparity between /f-plane and//-plane
patterns may be a disadvantage.
Such increased length elements are clearly impracticable for frequencies as low
as 30MHz, and attempts have been made to reduce the element lengths by end
loading. Difonzo34 gives a number of examples, the most significant of which are
shown in Fig. 17.58. It is also possible to add series loading in the form of a coaxial
line as shown in Fig. 17.59 and at lower frequencies a series coil has been used.
Although the discussion to date has been entirely on planar types, these are by
no means the only type of log periodic. The more general pyramidal form consists,
as Fig. 17.60 shows, of two sets of elements in inclined planes. By variation of the
angle between the planes, the E- and H-plane radiation patterns can be made more
nearly equal than with a single plane array. More variety in elements is possible and
the three most used forms are shown in Fig. 17.61. These are known as the
trapezoidal, toothed and zig-zag types.
770 VHF and UHF antennas

17.3.4 Helical antenna


A single conductor wound as a helix can be regarded as intermediate between the
monopole and the loop. On the one hand, as the diameter/) (Fig. 17.62) is decreased
the antenna tends to a monopole, whilst as the spacing S is reduced to zero, the
antenna becomes a loop. It is therefore not surprising to find that the helical
antenna has a number of radiation modes dependent on D and S in terms of wave-
length.

disc tee

Fig. 17.58 Forms of LPDA end-loading

in

I °

plane of
short circuit
b
Fig. 17.59 Series loading o f log-periodic element
(a) Internal (b) External

For directive antennas only the axial mode is of importance. This arises when
the circumference C of the helix is approximately one wavelength. As a general
guide, C lies within the range 0.75A< C< 1.33X for the axial mode. The on-axis
radiation is essentially circularly polarised provided that the helix has more than a
few turns. Degradation with the helix of few turns arises because of the wave
VHP and UHF antennas 771

reflected from the free end which produces opposite-hand circular polarisation on
reflection at the ground plane.
The characteristics of the axial mode helix are summarized below:
_ / . 90
Radiation pattern E = sin cos 0
n ]J sin 0///2)

Beam width (half power) B = degrees


CKy/nSx
115
Beam width (first nulls) B = degrees

Directivity £> = 15Cfrtfx


Input resistance /? = 140Cx
2n+ 1
Axial ratio i4rt = —

Fig. 17.60 Pyramidal log-periodic array with trapezoidal elements

where n = number of turns of helix, Cx = circumference in free-space wavelengths,


Sx = spacing between turns in free space wavelengths, a = pitch angle, 0 = angle
with respect to helix axis and \jj = phase shift per turn (see Kraus).
These data refer to helices with the following dimensions:
:i5° 3/4<Cx<4/3 n>3
772 VHF and UHF antennas

Practically the helical antenna requires a dielectric support structure which must be
minimised to prevent modification of the phase velocity along the helix which
would alter the electrical characteristics. Two possible arrangements are shown in
Fig. 17.63. The ground plane should be at least X/2 in diameter. The diameter of
the conductor does not appear to be critical between 0.05 X and 0.006X. The
arrangement of the feed portion from the ground plane appears to affect both
impedance and axial ratio. If the feed point at the ground plane is on the helix axis,
the feed wire should have the same angle as the pitch angle (Fig. 17.64).
A full analysis of the axial-mode single-conductor helical antenna is given in
Kraus53 Chapter 7, and further details are in Chapter 7 of Vol. I.

zigzag

toothed

Fig. 17.61 Forms o f elem en t

A short multi-element helix is described by Kilgus.45 This has a broad radi-


ation pattern and circular polarisation, but is of narrow bandwidth. Quadrafilar
helices have been studied by a number of authors, the position being best sum-
marised by Adams et al.1 It appears that the axial mode can be sustained over
the range 0.4 < C\ < 2.0, which offers an appreciable reduction in frequency for a
given size, compared with the unifilar helix, albeit with a more complex feed sys-
tem. The four helices are fed progressively with 0°, 90°, 180° and 270° phase
shifts, which can be provided with baluns and couplers as may be most convenient.

17.3.5 Slotted cylinder


The small-diameter cylinder with axial slot has by itself a broad horizontal radi-
ation pattern of width 150° to 180° between half power points for D/X between
0.2 and 0.6. It is, however, possible to modify these patterns to obtain narrower
VHF and UHF antennas 773

beams and better front-to-back ratios. Two methods have been applied to the pylon
slot antenna. In the first, two parallel axial slots are cut in the cylinder and fed in
phase. A spacing of 0.4X in a cylinder of D/X = 0.6 produced a 3 dB beam of 70°
width and a front-to-back ratio of 12 dB.

C -TTD

Fig. 17.62 Helical antenna

dielectric supports
(typical )

Fig. 17.63 Axial-mode helix

For smaller diameter cylinders the addition of dipole elements to the edges of a
single slot has proved effective. The proportion of power in slot and dipole is
altered by adjusting the length and position of the dipole elements. An optimum
arrangement in a cylinder of D/X = 0.1 gives a beamwidth of 110° and a front-to-
back ratio of 10 dB. Operation over the frequency band 87-100MHz with a VSWR
of 1.3 has been reported (Bosse7).

17.3.6 Comb(nations of elements


The radiation pattern requirements of some television and broadcast transmitting
stations are irregular in shape to suit their intended service areas. Such requirements
774 VHF and UHF antennas

cannot be satisfied by a single conventional antenna, but need a combination of


elements whose position, amplitude and phase are adjusted to produce the required
pattern.

Fig. 17.64 Detail of feed

B
33.

Fig. 17.65 Shaped patterns of orthogonal dipole arrays


In phase
100° shift
Relative amplitude A: B = 1:8

The elements themselves can be any of the directional antennas described above.
Directional elements are preferred to omnidirectional ones because they are less
affected by support structures and the mutual coupling between adjacent elements
can be neglected. In calculating the pattern of an array of elements around a mast,
for example, it is essential to know the phase centre of each element. It may be
VHF and UHF antennas 775

necessary to measure this for elements such as Yagis or log periodic antennas for
which it is not readily calculated. Fig. 17.65 shows patterns for arrays of full-wave
dipoles plus plane reflectors on two sides of a mast, for various phase shifts between
the arrays.

17.3.7 Significant design parameters


In choosing directional antennas for a particular application a number of electrical
and mechanical parameters have to be considered. The weighting to be applied to
each factor will, of course, depend upon the application, but a typical list would
include the following factors:
Electrical
Gain
VSWR
Coverage
Power handling
Bandwidth
Side-lobe level
Back-to-front ratio
Cross polarisation
Mechanical
Weight
Dimensions
Wind loading
Weatherproofing
A few examples may serve to indicate the importance of some of the less obvious
factors:
(i) Point-to-point links: In a chain of relay stations the most important factor
may well be gain since coverage is only required along a line. Because of frequency-
planning restrictions, limitation of signal in unwanted directions is important to
avoid mutual interference, and hence back-to-front ratio, side-lobe level and cross
polarisation will be significant. Poor performance in these areas will mean that the
separation between systems using the same frequencies has to be increased. These
factors may also be important where several antennas share a common mast to
minimise breakthrough of transmitters into co-sited receiving antennas,
(ii) Navigation systems: The glideslope components of current instrument landing
system (ILS) ground stations in the 330 MHz band use antennas at different heights
on a common mast to form, with their images in the ground, overlapping beams in
elevation. High wind-loading of the antennas necessitates additional stiffening of
the masts to prevent differential movement of the antennas causing a significant
shift of the equisignal path. If the antenna performance can be altered by snow or
ice-loading, see Section 17.3.2 on the Yagi-Uda antenna, then electrical de-icing
may be required or some form of radome which would increase wind loading.
776 VHF and UHF antennas

17.4 Vehicular antennas

Many communication and navigation systems used on vehicles operate within the
VHF or UHF ranges since the propagation characteristics permit the use of low-gain
omnidirectional antennas physically well suited to the vehicle constraints. The term
'vehicle' here applies to any means of transportation and includes the following:
Aircraft
Wheeled and tracked vehicles — iand mobile'
Marine vessels
Railway trains
Guided missiles
Rockets
Space vehicles and satellites
Hovercraft
Pedestrians and animals
In all instances the performance of the antenna will be significantly modified by
the vehicle on which it is mounted. Whilst the antenna itself may be simple in elec-
trical design, the combination of antenna and vehicle constitutes an electrically
complex structure. An insight into the influence of the vehicle can be obtained
by consideration of antennas mounted on or adjacent to certain regularly-shaped
conductors — flat sheets, cylinders and cubes are the most relevant.
Whilst analytical methods exist for some of the cases - the circular sheet, and
the circular cylinder — recourse has to be made to numerical methods such as wire
grid modelling or GTD for less regular shapes.

11.4.1 Antenna siting methods


Providing an antenna for a specific vehicle is very much an iterative process. On the
one hand there are the electrical requirements: frequency range, bandwidth, VSWR,
gain and radiation pattern coverage, whilst on the other hand there are constraints
due to the vehicle. These might include the following:
Height
Weight
Size
Air resistance
Conspicuousness
Environmental conditions
Temperature
Pressure
Humidity
Vibration
Shock
Corrosion
VHF and UHF antennas 777

It follows, therefore, that only in those systems where the electrical require-
ments are least demanding will it be possible to provide a single antenna design
suitable for a wide range of vehicles. The choice of antenna may have an impact on
the structure or the performance of the vehicle, and it is therefore highly desirable
to predict the antenna performance at an early stage of the design both of the
antenna and the vehicle.
The general procedure should be as follows:
(i) Definition of shape of vehicle
Definition of radio installations required ('radio fit')
Definition of operational constraints
(ii) Formulation of required antenna coverage from radio system and operational
requirements
(iii) Study of possible antennas to satisfy (ii).
(iv) Construction of antenna installation diagram showing preferred arrangement
of antennas
(v) Submission to vehicle constructor or user for approval
(vi) Modification of (iv) until agreement reached with user
(vii) (a) Numerical analysis of antenna performance or (b) scale modelling
(viii) Engineering of chosen antennas if no suitable existing types
(ix) Full-scale trials with engineered antennas
It will be noted that up to step (vii) the basic antenna characteristics are assumed
and it is only after this stage has been completed that any necessary engineering is
undertaken. There will, of course, be exceptions when more basic investigations are
necessary because existing knowledge is inadequate.

Mathematical modelling: Electromagnetic radiation problems can always be


represented by an integral expression based on Maxwell's equations. Such repre-
sentations in the past have often proved academic since they could not readily be
solved and much effort was expended in analytically manipulating the expressions
into a form which minimized subsequent computational effort.
With the advent of high-speed digital computers much of the analytical and
repetitive effort required in the solution has been drastically reduced. As a result,
efforts have now been channelled into aspects of redefining the various repre-
sentations such that they become solvable using numerically orientated techniques.
Several computerised mathematical methods therefore now exist which enable
the radiated field and other antenna parameters satisfying Maxwell's equations to
be formed in various spatial regions surrounding the antenna and structure con-
figuration. The input to a suite of computer programs performing such an analysis
would be the specification of the structure's shape and the output would be a
printed contour for discrete directions in space of the radiated power at specified
frequencies and other parameters.
Two such numerical methods which have received considerable attention in
recent years are those associated with wire grid modelling and ray-tracing techniques.
778 VHF and UHF antennas

Wire-grid modelling: The unifying concept in this particular treatment of radi-


ation problems is the method of moments described by Harrington.31'32 This
general approach to radiation problems is essentially a reduction of the associated
integral equation to a system of linear algebraic equations in N unknowns, where
the TV unknowns are usually coefficients in some appropriate expansion of the
current over specific regions of the antenna and its supporting structure. The
resulting matrix equation is then solved by known techniques to evaluate the
current distribution over the whole configuration.
In the present context there are two possible segmentation schemes for describ-
ing the various regions of the antenna configuration: surface patches and a wire grid.
In the former scheme the surface of the structure is divided into a small number
of patches each small in wavelength, but the application of this method requires a
structure whose surfaces are either flat or have continuous curvature. Thus this
scheme may introduce errors when applied to structures with sharp discontinuities
such as aircraft wings, etc.
The wire-grid scheme utilises a wire grid to describe the contour of the antenna
configuration and is based on the principle that a wire mesh is equivalent to a con-
ducting sheet. This is intuitively acceptable since it is generally realised that there
is little difference between the radiation performance of the solid dish antenna
and one constructed from wire gauze. The difference in electromagnetic properties
of a solid body and its wire-grid replica may therefore be made arbitrarily small as
the wire-grid mesh size is decreased. The wire-grid modelling scheme also allows
more flexibility than that using surface patches and as such has been more widely
used for various radiation problems. Of the many references on this subject
Richmond,75 Harrington and Mautz,32 Miller et al.,66 Kubina and Pavlasek,56 and
Imbriale34 are most relevant.
The wire-grid modelling procedure has the capability of representing a structure
of arbitrary shape, and the only real practical limitation of the technique is the size
of the matrix equation that one is able to operate on. The matrix size is essentially
proportional to the surface area (in square wavelengths) that is to be modelled, and
hence it is generally recognised that such techniques are restricted to structures
whose maximum linear dimension is not large compared to the wavelength of
excitation.
Typically the length of each segment should not exceed 0.1X. If it is desired to
model the impedance of the installed antenna it may be necessary to use a more
dense grid close to the feed point because of the rapid change of current across the
surface. The diameter of the wires is important, the aim being to match the
inductance of the grid to that of the continuous surface. Since a close mesh is
wasteful in number of segments, the use of thick wires (i.e. wires whose diameter
equals the minimum segment spacing between the centres of segments) is preferred.
Wire-grid modelling has been used extensively for VHF antennas on satellites,
but has been applied to low-band VHF on land vehicles. Fig. 17.66 shows measured
and calculated azimuth patterns for a low profile antenna on a tracked vehicle
approximately 0.8X long, 0.4X wide and 0.3X high.
VHF and UHF antennas 779

It should be noted that the use of this technique for calculating radiation
resistance does require care, particularly with short antennas. Small errors in the
phase of currents close to the feed point can lead to large errors of magnitude in
the computed radiation resistance.

340

30

280

260° ioo°

200° 160°
180°

Fig. 17.66 Comparison of predicted and measured patterns on tracked vehicle


X/4 monopole at 35 MHz on nearside top front of 5 m long vehicle
Predicted EQ E^
Measured EQ —-
Scale 4 0 dB

Ray-tracing techniques: Ray-tracing techniques inherit their basic concepts


from geometrical optics (GO). In the classical theory, GO predicts electromagnetic-
wave propagation along ray paths (which are straight in a homogeneous, isotropic
medium), zero fields in the shadow region beyond a scattering obstacle, and that
reflected energy is assumed to emanate from a locally plane point on the scatterer.
The major disadvantage of GO is thus the inability of the method to match the
natural electromagnetic processes at tips, edges, shadow regions etc. of a scattering
obstacle.
The geometrical theory of diffraction (GTD) may be described as an extension
of classical GO to include diffracted rays (Keller44 and Kouyoumjian52) and allow
780 VHF and UHF antennas

for the wave nature of light. These rays emanate from points at infinite curvature
(edges, tips etc.), when present on the scatterer. Other diffracted waves (creeping
waves) are waves which travel along smoothly curved surfaces shedding diffracted
rays tangentially into the surrounding medium. This theory thus enables predictions
to be made for the partially illuminated region behind an obstacle.
GTD can be used to construct the distribution of field intensity, polarisation and
phase throughout space for an antenna mounted on a structure, by tracing rays of
energy from the antenna incident on the scattering structure and then to a distant
field point. Each ray is assigned a field value, the phase of the field of the ray is
assumed proportional to the optical length of the ray from a datum and the
amplitude is assumed to vary in accordance with the energy-conservation principle.
Since GTD is based on an asymptotic high-frequency solution of Maxwell's
equations and the waves incident, reflected and diffracted on or from a scatterer
are assumed plane, this ray-tracing technique is applicable to, and has been used on,
structures which are large compared to the wavelength of excitation.

Fig. 17.67 Selection of rays from an antenna at P used in pattern predicted by G TD


A: Direct F: Pod creeping wave
B: Wing reflected G: Pod front diffracted
C: Wing edge diffracted H: Tail plane reflected
D: Nose diffracted I: Tail fin reflected
E: Fuselage creeping wave J: Creeping wave and reflection

Fig. 17.67 shows a selection of the rays which have to be considered on a com-
plex structure such as an aircraft. It should be noted that these modelling techniques
still have some difficulty in the axial direction of cylinders. This arises from the
importance of the end conditions and the increased ray complexity required for
accurate modelling.
VHF and UHF antennas 781

Scale modelling: Scale-modelling techniques described in this Section are con-


cerned only with the definition of scale-model vehicles for antenna siting. Clearly
with a vehicle of complex shape such as an aircraft the construction cost will
depend very much on the accuracy and detail incorporated. Some general rules are
given below; these are taken from IEC draft Publication 489 Part 2, 'Methods of
Measurement of Antennas for the Mobile Service'.
'The scale model shall represent the real vehicle in terms of:
(a) Conductive and non-conductive surfaces
(b) Discontinuities between conductive surfaces
(c) Fixed external projections greater than 0.1X in any linear dimension
(d) Variable external projections the position and shape of which are likely to have
a significant effect on the radiation pattern
(e) Surface roughness
( / ) Conductive structures, covered by non-conductive surfaces the linear dimen-
sions of which exceed 0.1 A
(g) Conductivity of skin'
Where the vehicle has a continuous conducting surface, only the representation
of that surface is important. Problems arise when there are apertures in the skin
since the size of the conductive cavity behind each aperture may modify the surface
current distribution. In an extreme case such as a small helicopter the amount of
metal around the cockpit may provide negligible screening. It is then necessary to
model all the internal metal structure whose size conforms to the criterion given
above.
Radiation patterns are very largely influenced by the current flow over the
vehicle surface, so any discontinuities have to be modelled. This can require very
detailed examination of the construction of the actual vehicle. A number of com-
mon hazards are given below:
(a) Gap between fixed and moving surface acts as a slot; length defined by distance
between metallic hinges
(b) As (a) but in aerofoil surfaces; the distance from a hinge to the open edge of
the aerofoil acts as a parasitic notch
(c) Removable panels isolated from structure by plastic or rubber strips and fixed
by a few bolts; distances between bolts act as slots
Considerable detective work may be necessary to determine where earthing
points exist as it is not unusual for hinges to have PTFE bushes, the actual earthing
being by a long wire braid. On some small military aircraft, contact between the
complete wing and the fuselage may be completely absent at the leading and trail-
ing edges and limited to fixings at the spars.

17.4.2 Influence of finite conducting surfaces


Distinction must be made between two different effects associated with the finite
metal surface:
782 VHF and UHF antennas

(a) As reflector or scatterer


(b) As ground plane
In the more general vehicular case, the second effect is the more usual. Two
aspects have to be considered:
Impedance
Radiation pattern and gain

-5

d
8 -1°
c
o
u
S -15

l-ao

-25-

1.0 a/A 2.0

Fig. 17.68 Impedance of 0.224X monopole 0.006X diameter on circular ground plane of
radius a
• • • • Meier and Summers results

Impedance: Meier and Summers64 show incremental changes to the impedance


of monopoles of various electrical lengths mounted at the centre of square and
circular flat ground planes. Unfortunately their paper does not give results for
planes smaller than 2X diameter, for which theoretical results are shown (Fig.
17.68) by Awadalla and Maclean.3 The variations of impedance arise from the
mutual coupling of the ground plane edges to the driven element — in a circularly
symmetric arrangement all the edge contributions have the same phase whereas the
phase variations in an asymmetric mounting give some impedance compensation.
VHF and UHF antennas 783

The current flow on a cylinder from a monopole mounted on it clearly differs


from that on a flat ground plane. It should not be surprising to find that a mono-
pole mounted on a circular cylinder 0.05X in diameter and X long, at mid-length,
shows negligible change in impedance as the cylinder length is increased. Similar
results apply for square, or rectangular, cross-section but with aerofoil sections
the variations approximate to those of a flat sheet.
A monopole mounted on the axis of a conducting element forms with the
element an asymmetric dipole. The impedance of such a dipole is given by the
mean-value theorem of King.46 This considers the asymmetric dipole to consist
of two monopoles of lengths L\ and L2 having impedances Zx and Z 2 , respect-
ively, where 2ZX and 2Z 2 are the impedances of the centre-fed dipoles of lengths
2LX and 2L2. Then the approximate impedance of the asymmetric dipole isZi 4- Z2.

15 20 25 30 35 40 45 50 60

Fig. 17.69 Impedance of \ / 4 monopole on car


A n ten na p osi tion
CR: Centre of roof
RR: Rear of roof
FR: Front of roof
FWF: Front of front wing
FWA: Rear of front wing
RWF: Front of rear wing
RWA: Rear of rear wing
150
—X— Frequency, MHz

In the case of a non-circular conductor an equivalent circular radiator can be


calculated, and from the published curves, e.g. Brown and Woodward,10 its
impedance can be determined. On a vehicle of complex shape and considerable
length it may be difficult to decide the appropriate length for the conductor pro-
vided the equivalent circular radiator is larger than 0.05 X diameter it will be suf-
ficiently accurate to assume a length of 0.5 X, since any length greater than this
will produce little change in impedance.
784 VHF and UHF antennas

It should be clear from this discussion that a monopole mounted, respectively,


on a flat sheet, a cylinder, or on the end of a conductor (e.g. aircraft tail fin), may
have a significantly different impedance in each case.

9=0°

yaw
plane
pitch
plane
0=180°

roll plane

Fig. 17.70 Monopole on cylinder: definition of axes

The practice of designing antennas on flat ground planes when they are intended
for use on three-dimensional bodies has little to recommend it except cost. It makes
much more sense to use a metal cylinder or cube of size appropriate to the ultimate
vehicle. As indicated above, if the cross-section is sufficiently large the length of the
body need not exceed one wavelength. This choice of a proper ground plane for
antenna development can be important in the determination of performance
characteristics. For example, a broad-band VHF aircraft antenna gave VSWR figures
of 4 at the band-ends against a specification of 2.5, but investigation showed that
the impedance had been measured on a 1.8 m square ground plane whilst the
antenna had been developed on a 2.4 m-diameter aircraft fuselage, on which the
specification figures were consistently met.
Further emphasis to the variation of impedance on a finite body is given by Fig.
17.69 which shows base impedance of a VHF monopole for seven positions on a
saloon car.
VHF and UHF antennas 785

Radiation pattern: In most instances the antenna is mounted on the vehicle, so


the dominant effect will be due to its three-dimensional structure. The simplest case
is that of a plain cylinder such as the body of a missile, rocket or satellite launch
vehicle, neglecting any guidance fins. The antenna may be classified either as a low-
angle radiator (monopole) or normal radiator (slot or parallel dipole) depending on
its main direction of radiation referred to the cylinder axis.

bottom

Fig. 17.71 Roll-plane pattern of centre roof VHF monopole on car. Free-space pattern

Referring to Fig. 17.70, computation of the radiation pattern in the yz (roll)


plane can be done analytically if the antenna is one of the basic types and can be
considered as a single point radiator. An array around a cylinder has to be treated
as the superposition of the patterns of its individual elements taking account of
their phases and amplitudes. A characteristic free space roll plane pattern for a
VHF monopole on a car roof (Fig. 17.71) shows that the pattern is dependent on
the whole size of the car and not on the roof size alone.
Patterns in the xy (yaw) and xz (pitch) planes are dependent on the radiation
from the ends of the cylinder, and hence on the form of those ends. Mathematical
modelling is difficult when the cylinder ends cannot be approximated by simple
geometric shapes, so it is not surprising that at present calculated patterns in these
planes are least accurate in the vicinity of the cylinder axis. This constraint only
applies when the cylinder diameter is electrically large, greater than about 0.1 A.
It should be obvious that close proximity to the cylinder end will produce highly
modified patterns.
786 VHF and UHF antennas

Pattern dependence is not limited to the end directions as Fig. 17.72 indicates.
This is a pitch plane pattern of a longitudinal horizontal dipole below a cylinder 4 A
long and 1.0X diameter, for different positions of the dipole. The strong radiation

scale in DB.

below

antenna position
— 1.15 A from nose
1.91 A "
2.67 A "

• AA

nose i

""*—*~ below

Fig. 17.72 Pitch plane patterns of X/2 dipole parallel to cylinder

in the nose and tail directions (end-on to the dipole axis) should be noted; this
would not occur if the dipole were mounted below a plane reflector and is a conse-
quence of skew currents on the cylinder. If the antenna is mounted on the cylinder
axis at one end of it, the combined structure may be considered as an asymmetric
VHF and UHF antennas 787

dipole whose patterns (see Jasik38 p. 3.16) depend on the relative lengths of the
two arms. The exact geometry of the adjacent end will have a marked effect if the
cylinder diameter is electrically large.

17.43 Vehicular constrain ts


The design of antennas for vehicular use is subject to a number of constraints not
encountered In other fields of antenna design. It is these constraints which have
prompted most of the variations from classical antenna forms. Three classes of
constraint will be considered here:
(a) Physical
(b) Mechanical
(c) Environmental

(a) Physical: These impose limitations on the height and size of antennas for
any of the following reasons:
Ground clearance, e.g. underneath aircraft
Overhead clearance: road vehicles, trains, helicopters
Hazards to vehicle operators and passengers
Interference with other vehicle functions
Launch-vehicle constraints: satellites, missiles
Concealment of vehicle
Disguise: plain vehicles for law enforcement

(b) Mechanical
Aerodynamic drag
Wind loading
Ice loading
Wave force: ships, submarines
Damage by cleaning equipment: cars, trains, public-transport vehicles
Weight: aerospace vehicles

(c) Environmental
Vibration
Shock, e.g. satellite launch, gunfire
High temperature
Low temperature
Effect of ice on electrical performance
Low pressure causing flashover and corona
Contamination by fuels and lubricants
Some of these constraints will be obvious on reflection, but others arise from the
possibility of consequential damage. For example, an antenna may be made
788 VHF and UHF antennas

sufficiently strong to withstand ice build-up in flight on an aircraft, but ice breaking
off the antenna may cause damage to engines and other parts of the structure.
Similarly the breakage of an antenna through vibration not only detracts from the
radio-system performance but the broken element may lodge in a moving control
surface and hazard the aircraft safety. It may also cause damage on the ground if it
falls clear of the aircraft.
The effect of low pressure is clearly seen from Fig. 17.73 in which the power
handling of a specific antenna is related to air pressure. The very low powers which
cannot be exceeded without breakdown represent a very real constraint on system
performance of high-altitude vehicles.

100

50

a
% 20

c 10
O

1
k
-L. J- -L. J
0.01 0.02 0.05 0.1 0.2 0.5 1 2 5 10
air pressure,mmHg

Fig. 17.73 Typical power breakdown as a function of air pressure. XIA bent monopole

Fig. 17.74 Simplified model of antenna on wheeled vehicle

17.4.4 Land-based vehicles

(a) Cars and vans: In Section 17.4.2 it was shown that the roll plane pattern of
an antenna on a saloon car, measured in free space, indicated that contributions
from the whole vehicle had to be included. It must be concluded that, when the
vehicle is on the ground, currents will flow into the ground through the wheels, or
VHF and UHF antennas 789

tracks in the case of a tracked vehicle. The impedance to ground through these
paths will, therefore, be important in determining the radiation patterns of an
antenna on a land-based vehicle, and in some circumstances may also affect the
feed-point impedance of the antenna. A simplified model of the system might be
Fig. 17.74 which suggests that the vehicle might be considered as a sleeve mono-
pole or, if the base impedances are high, as an asymmetric dipole.
Clearly the influence of the ground will decrease as the vehicle becomes elec-
trically large. Webster90 indicates that in the 30-76 MHz band currents throughout
the body and the wheels of military vehicles were significant. Some workers have
claimed that on motor cars effects can still be seen at 150 MHz with antennas on
the upper surface, but there is little published data. At frequencies of this order,
the capacitance to ground of the lower surface of a car must result in a low
impedance between vehicle and ground, so that the model is simplified further by
short-circuiting Z in Fig. 17.74. For practical purposes the effect of varying ground
constants can be neglected above about 100 MHz.
For terrestrial communications systems only a narrow range of angles above the
horizon is significant. Davidson23 shows that in a heavily built-up area propagation
into streets from an elevated base station may involve elevation angles up to 20°,
the dominant mode being over the roofs and by diffraction into the streets below.
Since most mobile antennas have broad elevation patterns, this effect only becomes
of importance in higher UHF bands where some elevation gain would be possible
through vertical stacking.
Again, it has been shown by Lee59 and Stidham86 that severe fading in cities can
be reduced by the use of antennas with horizontal directivity, provided that the
beam maximum is pointed towards the base station. For a mobile whose direction
is constantly changing this suggests that some form of adaptive antenna would be
desirable, but, except for special cases such as the need to maintain continuous
communication with high data rates, the complexity of such systems is rarely
warranted. These results do suggest, however, that for purely city use an insistence
on very good omnidirectional coverage is unnecessary. This does not apply in more
open terrain; field trials in town and country with good and poor horizontal
coverage antennas clearly show the need for good omnidirectional coverage for
rural use.
Measurements on cars in the 80MHz, 150MHz, 450 MHz and 900 MHz bands
show that best coverage is obtained with a centre roof antenna. A typical pattern is
shown in Fig. 17.75 which should be contrasted with Figs. 17.76 and 17.77 for
front and rear roof positions. It should be noted that the front position has maxi-
mum signal to the rear and vice versa; pitch plane patterns such as Fig. 17.78 show
considerable distortion and help to explain this paradox.
Wing mounting gives more distorted patterns as might be anticipated (Fig.
17.79), and the mean gain is about 3 dB below that for roof positions.
Measurements on delivery vans show less variation in pattern with roof position.
The radiation patterns for roof-mounted antennas on cars do not correspond to
those on flat ground planes of the same size. Strong currents can be measured in the
790 VHF and UHF antennas

roof pillars, and these clearly influence the patterns and explain why the patterns
on metal-sided vans are less sensitive to antenna position.

Antenna types: Antennas for use on cars and vans are usually X/4 whips of very
thin (1.5 mm diameter) spring steel. 5X/8 or 3 X/4 whips can be used in the UHF
band; these have a loading coil at about X/4 above base to give cophased currents
along the whip. As noted by Davidson and Turner,23 the full gain of such antennas
may not be realisable in built-up areas.

front

Fig. 17.75 Yaw-plane pattern of VHF monopole on centre roof of car

Whilst the centre-roof position is clearly the best for electrical performance
some mobile operators prefer not to drill holes in car roofs, and fit antennas to the
wings instead. Some use has been made of wing-mirror antennas which combine
two functions. Essentially these are short monopoles, insulated from the car body,
and incorporating a preset matching section in their bases.
Although their very low cost makes whip antennas attractive for the civil user,
they are vulnerable to damage by several causes:
Vandalism
Low overhead clearances: bridges, garages
Automatic car washes
VHF and UHF antennas 791

Alternative low-height antennas that can be used are described in Section 17.2.1
under the heading 'Vertically polarised monopoles'. The 'low profile' (Fig. 17.19)
and annular slot (Fig. 17.20) have been widely used for public-service vehicles.

front

Fig. 17.76 Yaw-plane pattern of VHF monopole on front roof of car

(b) Other wheeled civil vehicles: Vehicles which do not have continuous metal
upper surfaces present special problems. Examples include fire appliances, flat-bed
lorries, earth-moving machinery, tractors and articulated vehicles. Each of these has
to be considered separately. The increased height of some gives a height-gain which
compensates for the irregular radiation patterns. Measurements on a whip antenna
on a fire appliance at about 80 MHz showed a mean gain of at least 6 dB over a
similar roof-mounted antenna on a saloon car.
On some vehicles it may be more appropriate to use some form of dipole in the
absence of a well-defined groundplane. The coaxial dipole Fig. 17.9 has commonly
been used in such applications.

(c) Military vehicles: Most communications systems for military vehicles


operate in the range 30-100MHz. Whereas civil systems can use an elevated base
station with relay facilities for car-to-car operation, this is not possible for military
tactical communications and the lower propagation losses favour the use of the
lower VHF bands.
792 VHF and UHF antennas

Antennas are normally X/4 whips with spring-loaded bases to minimise damage
from obstacles. The need to match the antenna at any frequency within the band,
typically 30-76 MHz, demands the use of complex tuning units which can accom-
modate the wide range of impedances encountered on a variety of vehicles, none of
which provide a large ground plane. Coaxial dipoles with tuned decoupling (Fig.
17.10) are commonly used; their impedances and radiation patterns are less
sensitive to the shape of the vehicle.

Fig. 17.77 Yaw-plane pattern of VHF monopole on rear roof of car

Since it is common to fit two or more radio systems to the same vehicle, the
provision of additional selectivity, either in the tuning unit or by separate tunable
filters, is common practice to reduce the frequency separation between equipments.
Whip antennas, typically 1.8 or 2.5 m long, have a number of disadvantages:
Vulnerability to damage by obstacles, gunfire or explosives. Visibility when the
remainder of the vehicle is concealed.
Various low-profile antennas have been considered. These are generally of the
top-loaded monopole form in which the top-loading is a metal plate on a short
vertical stub, the whole antenna being supported by a reinforced plastic fairing.
Such antennas have very low radiation resistances, and since they are essentially
capacitive their tuning-unit losses are high and the antenna efficiency may be at
least lOdB below that of a tuned monopole. One variant of Fig. 17.19 which
VHFand UHF antennas 793

appears to combine the merits of robustness, low height and good efficiency is
shown in Fig. 17.80. The conventional form can be tuned over the whole frequency
band but only at the expense of low azimuth gain at low frequencies where a large
capacity loading is necessary. By shorting out parts of the top loading coil the range
of capacitance is reduced and the azimuth gain does not fall below — 5 dB com-
pared with a tuned monopole. The whole structure is supported within a thick
GRP cover having an overall height of less than 0.5 m. The relatively high Q of this
antenna enables two separate equipments on the same vehicle to be operated with
frequency separations of no more than 12%.

\
\

front

bottom
Fig. 17.78 Pitch-plane pattern of VHF monopole on front roof of car. Free-space pattern

17.4.5 Marine vehicles


The use of VHF and UHF on marine vehicles is limited to a narrow band around
160 MHz for ship-to-shore use and to 225-400 MHz for naval vessels. Effects of the
masts and superstructure vary so much from vessel to vessel that no general rules
can be given for siting except to mount the antennas as high as is practicable.
For single VHF installations some form of coaxial dipole Fig. 17.9 is usually
employed. Where more than one installation is required the antennas are carried on
outriggers to minimise the effect of the support structure. Section 17.2.3 gives
general guidance on the effects of such structures.
UHF antennas are normally a form of bicone, axially fed. Fig. 17.81 shows a
794 VHF and UHF antennas

typical form, the shape of the lower cone being modified to minimise currents on
the transmission line. A particular design feature is the use of large insulators not
only for strength but to reduce the effect of encrusted salt and funnel effluents on
the performance of the antenna.

Fig. 17.79 Yaw-plane patterns of VHF monopole on front and rear wings of car

17.4.6 Aircraft
This section encompasses antennas for helicopters and fixed-wing aircraft. The
antennas will be considered under the following headings relating to the radio sys-
tems to which they belong:
Communications
Identification and air-traffic control
Telemetry
Navigation
Homing

Coverage requirements: Simple power budget calculations for ground-to-air


systems will show that the depths of the minima in the radiation patterns which
will be encountered in practice will rarely upset system performance except at
extreme range. The required coverage is therefore related to the attitude changes
of the aircraft in normal flight rather than to the fact that it will be at high
VHF and UHF antennas 795
elevation angles to the ground station at short range. Aircraft can therefore be
classified conveniently as aerobatic and non-aerobatic. The latter class includes all
civil aircraft except those used for aerobatic competitions and a large number of
military aircraft, too. In general, only strike and fighter aircraft, high-performance
trainers and some of the more manoeuvrable light helicopters need to be con-
sidered as aerobatic.

antenna element

variable capacitor

Fig. 17.80 Low-profile VHF antenna, 30-76 MHz


H « 0.05\ 30

For the majority of aircraft, therefore, the coverage requirements will be limited
to the azimuth plane and to elevation angles limited by normal aircraft attitude
changes. Angles of 4- 5° to — 30° with respect to the aircraft horizontal axes should
be adequate for most non-aerobatic aircraft. Clearly larger angles must be considered
for more manoeuvrable aircraft and the specification modified to suit individual
requirements.
796 VHF and UHF antennas

Satellite-communication systems naturally require rather different coverage. For


normal horizontal flight the elevation requirements will be + 10° to + 90°, that is,
upper hemispherical cover above + 10°.

upper cone

insulator

-lower cone

17.81 UHF naval bicone

Communications (Ground-to-air and air-to-air): All the commonly-used systems


use vertical polarisation and operate within the range 30-400 MHz. Gain require-
ments are modest — a result of the practical constraints in siting - and a minimum
level of — lOdBi within the coverage requirements will normally be adequate. It
can be shown in most instances that this will permit good system performance
within radio line-of-sight and a large and uneconomic system gain increase is
required to give significant increase in range. This subject has been discussed in
depth by Gough.28 Transmitter powers do not exceed 50W CW, so problems of
corona and flashover are not significant.
It follows, therefore, that on metal aircraft simple monopole antennas properly
sited will be adequate and the types described in Section 17.2.1 under the heading
'Vertically polarised monopoles' can all be used provided that they are suitable
aerodynamically. Thus only types vii, x, xiii, xiv and xv have not found application
for aircraft communications either because of shape in the case of types x, xiii and
xv or because of bandwidth in the case of types vii and xiv. It should be noted here
that current radio systems operating above 100 MHz have no facilities for antenna
VHF and UHF antennas 797

tuning, so the narrowband but tunable systems are operationally unacceptable.


Between 30 and 100 MHz constraints of physical size have necessitated the use of
tunable antennas, which will be discussed below.
The azimuth radiation pattern of a monopole on a long horizontal cylinder is
substantially circular provided that the antenna is more than about X/2 from an
end. For frequencies above 100 MHz, therefore, top or bottom fuselage-mounted
antennas are satisfactory. Mounting on or near the centre line is desirable to avoid
tilting of the roll-plane pattern (Fig. 17.82). The pitch-plane pattern is similar to
that of a monopole on a flat ground plane (Fig. 17.83) and for a balanced pattern a
position in the middle third of the fuselage length is preferred.

above

300 MHz

17.82 Roll-plane patterns of vertical monopole under aircraft fuselage


On lower plane, 5° tilt to port
Offset 0.3 \ to port

The wings can be considered as an extended ground plane and their effects will
be a combination of reflection and diffraction, the former dominating if the
antenna is at significant height above or below the wing surface. To minimise these
effects fuselage positions forward or aft of the wings are preferable.
For upper fuselage positions, shadowing by the tail fin has to be considered. Fig.
17.84 due to Cary (unpublished) shows the effect of an elliptic cylinder for parallel
798 VHF and UHF antennas

and perpendicular polarisation. It is evident that on large aircraft the shadowing


may lead to unacceptable minima particularly at the higher frequencies. Fig. 17.85
shows a typical azimuth pattern of an upper fuselage antenna.

above

nose 4—tail

300 MHz

Fig. 17.83 Pitch plane patterns of vertical monopole under aircraft fuselage
Under forward fuselage
Under rear fuselage

Where it is impracticable to find a satisfactory site for a single antenna, it may be


necessary to use two antennas with some method of selection. Permanent connec-
tion of two widely spaced antennas is impracticable because of multiple interference
lobes and unequal power sharing arising from differing impedances at the junction.
In these circumstances it is usual to choose one forward-looking and one rearward-
looking position. Two wing positions could be used, but practical considerations of
cable attenuation and difficulty in running the cables will usually make this arrange-
ment less preferable. If wing positions have to be used, an antenna on the upper sur-
face of one wing should be complemented by one on the lower surface, both near
the wing tips to obtain the best elevation coverage.
An alternative position, much used on high-performance aircraft, is at the top of
the tailfin, the antenna being enclosed in a dielectric cap. Monopoles of types iii and
iv are commonly used. The subject is covered in some detail by Mahoney.61 On an
aircraft with a high tail plane it may be unnecessary to enclose the antenna in a cap
and a conventional fuselage type can be used. One bonus of the high tail plane is
VHFand UHF antennas 799

the limitation of currents down the leading and trailing edges of the tail fin which
will normally give rise to interference lobes. For VHF communications in the
100 MHz band this is a very good position. For higher frequencies the cable attenu-
ation may outweigh the advantages.

source of vertical
polarisation radiation thickness
vibrating parallel to
streamlined cylinder

Fig. 17.84 Effect of a streamlined obstacle on radiation pattern of a dipole parallel to its
own axis
(a) Horizontal polar diagram with streamlined cylinder of 0.1 X thickness, 1.0X
chord
(b) Horizontal polar diagram with streamlined cylinder of 1.OX thickness, 10.0X
chord
r = 10\

Frequencies between 30 and 100 MHz are mainly used for military communi-
cations with mobile ground forces, so tend to be required only for helicopters and
small aircraft. A maximum length of under 20 m means that at the lower frequen-
cies the fuselage is never more than 2A and may be significantly less. When, as is
often the case, the fuselage cross-section is non-uniform, unequal division of current
on the fuselage forward and behind the antenna results in distorted and very
frequency-dependent radiation patterns. Fig. 17.86 illustrates the rapid change in
yaw plane patterns on a centrally mounted helicopter antenna, described in more
detail by Burberry and Kelly.13
Apart from this problem of siting, there is a further difficulty in this frequency
band since wavelengths are too long to permit the general use of self-resonant
antennas. The alternatives are short, inductively-loaded monopoles such as that
described by Cooper19 and shown in Fig. 17.87, bent-sleeve monopoles with tuning
800 VHF and UHF antennas

units, or variants of the transmission line antenna (Fig. 17.19). Inevitably, gain at
the lower end of the frequency range is low compared with a A/4 resonant mono-
pole, probably about — 10 dB for the inductively loaded types and — 6dB for
capacitive loading. At the upper end of the band efficiencies approach that of the
monopole.
Quarter-wave whips have been used on the slower helicopters (speeds less than
about 150km/h) but have to be mounted at the tip of the tail fin or underneath the
fuselage with some means of retraction. Simple switched matching systems are
possible and gains are comparable with resonant monopoles.

starboard

nose 0

port

antenna

Fig. 17.85 Typical azimuth pattern o f upper fuselage antenna

Another possibility is to use the tail fin as the antenna, treating it as a folded
monopole or shunt antenna (Fig. 17.88). The radiation patterns are very much a
matter of chance and each installation must be treated on its merits. It seems prob-
able that successful operation will only be possible where (a) there is no rudder, (b)
the chord of the fin is less than A/4, and (c) the fin height is less than A/2. Some
form of tuning unit will be required.
The large frequency band, together with the use of antennas very much physic-
ally constrained, necessitates tuning systems. In the simplest type band-switching
may be possible, but in others tuning at each frequency is essential. If frequency
VHF and UHF antennas 801
information is available from the radio equipment some pre-setting of the tuning
elements can be achieved whilst the transmitter is tuning, final trimming if necessary
requiring RF power from the transmitter. This method does permit silent tuning for
reception. In the absence of frequency information, antenna tuning must have RF
power from the transmitter and this results in longer tuning times.

\
\

port \/\starboard

antenna

Fig. 17.86 Azimuth patterns of VHF antenna under asymmetrical fuselage


26 MHz
32MHz
38MHz
42MHz

Satellite communication: Whilst VHF and UHF systems for air-to-ground com-
munications via satellites do not require high antenna gains the need for circularly-
polarised hemispherical coverage cannot readily be achieved on an aircraft with a
single antenna because of physical constraints. Most systems have in fact been a
compromise between high-angle circular polarisation and low angle vertical
polarisation. A typical arrangement (Fig. 17.89) consists of a pair of crossed slots
with shallow cavities for the high-angle cover linked with a vertical monopole for
the lower angles.
802 VHF and UHF antennas

Identification and air-traffic control: The frequency range 960-1220MHz is


occupied by several radio systems used for identification (IFF), air traffic control
(ATC) and distance measuring (DME). These are all pulse equipments using on-
board transmitter powers of the order of 1 kW peak and using vertical polarisation.

Fig. 17.87 Schematic of inductive loaded monopole 30-76 MHz

dielectric fairing

Fig. 17.88 Tail-shunt antenna

The coverage requirements tend to be more stringent than for communications;


this is understandable since the penalties for loss of returns from an aircraft may be
more severe, especially in a military environment. Basic gain requirements can,
VHF and UHF antennas 803

however, still be met by monopole antennas, the small wavelengths permitting the
use of highly efficient tuned monopoles and of annular slots (Fig. 17.21), where
flush antennas are necessary. In this respect it should be noted that the weight of an
annular slot is of the order of 1 kg compared with perhaps 0.20 kg for a monopole
of good aerodynamic design, and there will rarely be an advantage in using the flush
antenna.

upper surface
of fuselage

Fig. 17.89 Slot-monopole combination for aircraft-satellite communications

port starboard

tail
Fig. 17.90 Typical yaw-plane pattern of 1 GHz vertical monopole centrally mounted under
large aircraft fuselage

DME and ATC systems essentially require coverage below the azimuth plane,
and a single antenna on the underside of the fuselage will normally suffice (Fig.
17.90). IFF requires some upward coverage for air-to-air interrogation as well as to
804 VHF and UHF antennas

allow for aircraft manoeuvre. The addition of a second antenna above the fuselage
is not adequate since the tail fin will give severe rearward shadowing. A better
arrangement is to have one antenna on the upper surface of the nose or even in the

Fig. 17.91 Typical positions for optimum IFF coverage

feed point of
driven notch
Fig. 17.92 Triple notch in tapered section
P: Parasitic or stop notch

nose, and a second one under the tail (Fig. 17.91). The positions of each have to be
chosen to make good the deficiencies of the other. If there are no suitable surfaces
for vertical monopoles, horizontal slots or notches can be used. Fig. 17.92 shows
VHF and UHF antennas 805
12
a notch antenna for mounting in the trailing edge of a tail fin (Burberry ). Parasitic
elements above and below the driven notch reduce current flow on the fin and give
azimuth radiation patterns similar to those of a vertical dipole and director.

Telemetry and command: Telemetry systems on aircraft are mainly used in the
flight test phase and command antennas are restricted to pilotless aircraft which
may be targets for weapon trials or RPV (remotely piloted vehicles). A variety of
frequencies have been used, but in general the coverage requirements have to take
account of large aircraft manoeuvres and will be similar to those for IFF. Most of
the systems are narrow-band, and can therefore use small antennas of monopole
types, the transmission-line antenna being frequently used in the 400MHz band.

Navigation: This Section refers exclusively to VHF instrument landing systems


(ILS) and VHF omni range (VOR), both of which are used for the navigation of
civil aircraft in controlled airspace. Both systems use horizontally polarised ground
transmitters, the on-board equipment being used solely for reception.
The ILS system has three components:
Localiser 108-112 MHz Azimuth guidance
Glidepath or glideslope 328-336 MHz Elevation guidance
Marker 75 MHz En route position

ILS localiser and glidepath: The localiser and glidepath equipments in the air-
craft compare the depth of modulation of two signals radiated from the ground
station. The equisignal path is defined as that on which equal modulation depths
are received, and hence the accuracy with which this path can be followed depends
on the ability of the aircraft antennas to reject false information. The signal
received at a point in space, neglecting the effects of the aircraft, may arrive from
several directions as well as directly from the ground station. The additional paths
occur from reflections from objects on the ground or from other aircraft. Since the
radiation from the ground station has a difference in depth of modulation (DDM)
which varies with angle, the additional signals will each have a different DDM and
random phase. The combination of these unwanted signals with the wanted signal
produces the so-called 'bends' in the localiser and glidepath beams. For a given
ground station the severity of the bends as seen by the aircraft will be a function of
the aircraft antenna. The polarisation of the unwanted signals may also be random,
so that the response of the aircraft antenna to the unwanted vertically polarised
component may also be significant.
A further source of error arises from the signals received at the antenna by
re-radiation from parts of the aircraft. The vector sum of the aircraft contributions
is a function of aircraft attitude, and as this alters, the apparent position of the air-
craft also moves with reference to the equisignal line. Moving parts of the aircraft —
propellers, helicopter rotors and radar scanners — may also impose further modu-
lations on the received signal. Whilst many instances have been recorded of the
806 VHF and UHF antennas

serious effect of rotors and propellers on both localiser and glidepath reception,
radar scanners have been observed only to affect glidepath signals because their
dimensions are small in terms of wavelength at localiser frequencies. Studies of
phase-centre shift of ILS localiser antennas have been carried out extensively by
the Royal Aircraft Establishment (RAE) in England (Jones41).
Gain requirments for localiser and glidepath antennas are modest, the usually
acceptable figures being — 19dB for localiser and — 15dB for glidepath, in the
forward direction, referred to a half-wave dipole. Coverage requirements vary
with the flight procedures adopted to acquire the beams, but in general these
minima should apply up to 45° in azimuth from the line of flight. Where localiser
antennas are also used for VOR reception, omnidirectional coverage will be
required and a rather higher gain because of the longer ranges involved.
The basic requirements are therefore:
(i) Adequate gain
(ii) Minimum phase centre shift with normal aircraft manoeuvre
(iii) Good polarisation discrimination
(iv) Directivity

ramshorn

nose dipole

pair of suppressed
or external aerials
pair of
suppressed aerials

Fig. 17.93 Positions for localiser aerials

Localiser antennas: Fig. 17.93 shows typical positions for localiser antennas.
Four main classes can be distinguished:
(i) Ramshorn
(ii) Pair of monopoles or flush elements
(iii) Wing tip or tail-plane tip antenna
(iv) Nose dipole
VHFand UHF antennas 807

Ramshorn: The ramshorn is a horizontal half-wave dipole bent to the shape of a


V or U and mounted approximately O.IX above or below the aircraft fuselage. On
propeller-driven aircraft it may suffer seriously from modulation, and in general
should be mounted well forward to minimise phase-centre shift. Maximum radi-
ation is normal to the fuselage surface, so azimuth gain is low and the discrimi-
nation against vertical polarisation poor.

Pair of elements: To improve azimuth gain the vee dipole was mounted through
the tail fin. This is unsatisfactory on thick fins and a pair of monopoles, raked 45°
and fed in anti-phase, is preferred. Half-loops and flush-mounted elements with
backing cavities are also used (Thomas and Johnson87). For high-performance air-
craft a notch-fed plate, similar to Fig. 17.28 and enclosed in a dielectric fairing, has
proved satisfactory.
This type of antenna on a fin should be mounted at least 0.5X below the top of
the fin metalwork to minimise edge currents. Because the system is electrically
balanced, vertically polarised radiation from the fin leading and trailing edges is
a minimum in the forward and rearward directions. If, however, the balance is
upset, there will be a significant increase in such radiation received by the antenna.
Fin-mounted antennas receive signals from the forward direction with strong
reflected components from the fuselage and wings so aircraft attitude causes
phase-centre shifts, which will be unacceptable for automatic landing systems
although normally not a problem for general ILS use.

radome'

balance unit

Fig. 17.94 Suppressed localiser aerial system in nose of aircraft

Elements may be mounted in each side of the aircraft nose if their separation is
not more than about 0.9 m. This gives good horizontal plane gain and has proved
satisfactory for ILS and VOR in spite of the inevitable grating lobes Fig. 17.94.
Space for flush antenna cavities will normally be limited and, since the dimensions
808 VHF and UHF antennas

need to be about 0.75 m long, 0.25 m high and 0.2 m deep for a system VSWR not
worse than 7:1 referred to 5012, there will only be limited application for this type
of antenna. Any of the external elements used for fin mounting can be used if
aero dynamic ally acceptable, types totally enclosed in dielectric covers being most
suitable and least prone to damage in ground handling.

Wing-tip antennas: Structurally the wing tip will often be suitable for a flush
monopole or capacitance-loaded notch. A single element is used. The leading edge
of the wing will have a strong current flow and this may lead to significant pick-up
of re-radiated signals from the fuselage, making the antenna response very
susceptible to aircraft attitude. Phase-centre shift is very marked, and this antenna
position is not suitable for automatic landing systems.

nose

port starboard

2 A in scanner

forward edge of metal

glidepath aerial

Fig. 17.95 Effect of weather radar scanner on radome-mounted glidepath aerial


— Scanner in central position
Scanner removed

Nose dipole: A dipole mounted in the nose radome is naturally shielded from
the remainder of the aircraft, and hence the dependence on aircraft attitude is a
minimum. Such an antenna has proved very satisfactory for automatic landing and
VHF and UHF antennas 809
for aircraft used for inspection and calibration of ILS ground stations. The antenna
is physically and electrically symmetrical, has adequate gain and good directivity.
It should be noted that correct disposition of lightning strips on the radome is
important to minimise pick-up of vertically polarised radiation and failures of such
strips may degrade the antenna performance.
It can be seen from the examples given that for the highest-quality service the
antennas must be mounted on the aircraft centre line and must be electrically and
physically symmetrical, with minimal reception of signals re-radiated by the aircraft.

Fig. 17.96 Bull-ring glidepath an tenna

Glidepath antennas: The criteria for localiser antennas apply equally to glide-
path antennas. Only limited forward and downward coverage is required and the
most suitable position is towards the nose of the aircraft. The movement of weather
radar scanners can seriously modulate the received signal as well as reduce the
signal level, as Fig. 17.95 demonstrates.
Since the bandwidth required is small, short antennas are possible. Variations on
the transmission-line antenna are commonly used (Fig. 17.96), the shape and load-
ing being adjusted to give maximum forward signal. This type can be mounted on
the aircraft nose, in a radome, or on the metal behind the radome.
As a flush-mounted antenna the half-wave slot with backing cavity is very suit-
able. Direct or probe-feeding may be used, the latter permitting two isolated feeds
810 VHF and UHF antennas

in the same cavity for multiple receiver installations. Typical cavity dimensions are
0.44 A long, 0.16A wide, 0.13 A deep, the slot with dielectric cover being only
0.40A. If the aircraft nose is occupied by a scanner, the slot can be cut in the metal
skin immediately aft of the radome, as shown in Fig. 17.97. With the latter
arrangement there may be significant reception from the rear, and propeller
modulation may be a problem.

20* minimum

Fig. 17.97 Positions for ILS glidepath slot antenna


(a) In metal nose
(b) Behind radome

Ideally the glidepath antenna should be on the pitch axis of the aircraft so that
pitch changes produce minimum change of antenna position. For this reason, tail-
fin mounting should not be used where the aircraft has a large pitch change in the
approach and landing phase.
On small and medium propeller-driven aircraft it can be difficult to find a
position free from modulation for the conventional antennas. Some success has
been achieved with a short Yagi array of director, driver and reflector in monopole
form mounted on the side of the aircraft nose (Fig. 17.98).

Marker antennas: Marker position-indication occurs when the signal at the air-
borne receiver exceeds a specified and preset level. For a given aircraft height the
received signal is dependent on the transmitter power and the product of the
ground and airborne antenna patterns. It can be shown that, because transmitter
powers and antenna patterns vary considerably, it is impracticable to satisfy the
specification in detail.
The ICAO standard antenna is a half-wave dipole spaced 0.038X below the air-
craft fuselage and with axis fore-and-aft. The position of such an antenna has a
significant effect on its radiation pattern as Fig. 17.99 indicates.
Because of its drag the standard antenna is rarely used, the alternatives being an
external or flush mounted transmission-line antenna (Fig. 17.19). In this appli-
cation, high capacitance-loading is necessary to obtain good radiation from the
VHF and UHF antennas 811

horizontal element as well as to achieve a physically small antenna. Typically the


element length is 0.06X. In the flush-mounted form the element lies in the open
face of a shallow cavity and this type of antenna is sometimes known as a bath-tub
antenna.

port starboard

azimuth polar diagram of aerial on port side of nose

folded monopole reflector


glass fibre
cover

director

Fig. 17.98 Narrow-beam glidepath aerial

VOR antennas: Provided that essentially omnidirectional coverage is obtained,


the localiser antennas described above will be suitable for VOR reception. The gain
requirements are comparable with those of a line-of-sight communications system, a
minimum azimuth level of — 8 dB referred to a half-wave dipole being adequate. Of
the localiser types the ramshorn and the pair of elements have proved the most
satisfactory, the latter being particularly good when fin-mounted external elements
are used.
It should be noted that, whereas phase-centre shift is normally only a significant
problem on the equisignal path of ILS, it may be important at any azimuth angle
on VOR. The cross-polarisation is low along the equisignal path from a symmetrical
and balanced antenna, but may be appreciable at other angles.
812 VHF and UHF antennas

Homing: The majority of airborne homing systems utilise ground stations used
for normal communications and are therefore vertically polarised. The principles
would apply equally to horizontally polarised systems, but these will not be
described here.

rtose

Fig. 17.99 Pitch-plane patterns of ILS marker antenna under aircraft fuselage
Length of fuselage: 2.5\
0.2\ from nose
1.25\ from nose

The aim of a homing system is to enable the aircraft to be flown towards the
transmitting station. If the apparent direction of the station given by the homing
system differs from the aircraft fore-and-aft axis the system is said to have a squint.
Homing to the station may still be possible but the flight path will then be curved.
The system must be free from ambiguities; if the ground station is to the left of the
aircraft, the homing system must show this. In summary the basic requirements are:
(i) Minimum squint
(ii) Good sensitivity on-course
(iii) Good angular discrimination
(iv) Freedom from ambiguities
VHF and UHF antennas 813

Two types of homing system can be distinguished:


Phase
Amplitude

aircraft
centre line

identical cables

phasing link

Fig. 17.100 Essen tial elemen ts of phase homing system

Phase homing: The essential components of a phase system are shown in Fig.
17.100. They consist of a pair of identical antennas symmetrically disposed about
the aircraft fore-and-aft centre line and connected by a phasing link. The take-off
point can be switched so that one antenna leads or lags on the other. Assuming for
the moment that the two antennas would individually receive identical voltage
amplitudes in the absence of the other, and that the mutual impedance between
them does not alter their relative phase, then the phase between them is
(2TTJ/X) sin 0, where 6 is the angle between the axis of the aircraft and the direction
of the beacon.
If the phase shift in the feed cable is 0, then the total shift is
\jj = (2nd/X) sin (6 + 0).
If Et and E2 are the voltages at each antenna and we write E2 = nE\ the resultant
voltage ER = Ety/l + n2 4- 2n cos \p.
As the take-off point is switched from A to B, \p will alter from (2TT^/X) sin (d + 0)
to (2nd'/X) sin (6 — 0). At any angle 0 the ratio ERA/ERB is the DF ratio. lfEx =E2i
ER - 2EX cos \p/2 and the DF ratio becomes (cos i//l/2)/(cos
814 VHF and UHF antennas

The basic equation for ER permits calculation of DF ratios for different values
of 0 using measured radiation patterns for Ex and^ 2 -
Ideally the DF ratio should increase monotonically as 6 increases from 0° to 90°.
Maximum ratio at 6 = 90° is achieved by making the value of \p at this angle equal
to it, a condition which has, in theory, to be met at the highest operating frequency
to avoid ambiguity. In practice the system can be arranged to limit at DF ratios
exceeding some set value and the restriction can be eased, provided a good site can
be obtained. Difficulties arise when the radiation patterns of the two antennas
differ to the extent that, say, the port antenna has a minimum on the port side
which is not matched by a similar minimum by the starboard antenna at the same
angle on the port side, As a general guide the spacing between antennas should be
A/4 at the highest frequency and <j> should be 90°. As the frequency is decreased
the phase shift decreases and the DF ratio changes more slowly, to the point that
the equisignal path is not clearly defined. It is normally reckoned that the DF ratio
at 10° off course should lie between 1.4 and 2.0 for satisfactory homing. Higher
ratios make it difficult to hold a steady course and are generally accompanied by
reduced signal level on the equisignal line — in the extreme case the on-course range
may be very much less than the off-course, making it necessary for the aircraft to
be yawed to re-acquire signal.
To achieve good homing accuracy, identity of impedance and of radiation pat-
tern are both important. It is not sufficient to have identity of VSWR; very close
control of impedance is essential and low VSWR will make this control easier. It
follows, therefore, that high homing accuracy over wide frequency bands is difficult.
Normal communications antennas, even of broad-band design, may not have suf-
ficiently identical impedance to give good homing. There is some evidence that
resistance loading may be advantageous. Since the left-right symmetry of radiation
pattern of an antenna mounted on an aircraft fuselage decreases with the displace-
ment of the antenna from the centre-line, it becomes increasingly difficult to obtain
good homing at low frequencies with this type of antenna system.
Antennas used for this type of homing will normally be A/4 monopoles or
possibly transmission-line antennas. The latter are useful only for single-frequency
applications where height is limited, as they are difficult to optimise. For operation
in the 30-76 MHz band, dipoles in front of the fuselage are the only form of
antenna suitable because of ground-clearance limitations and because the spacing of
approximately 1 m would give very asymmetric patterns for monopoles on the
average fuselage. It is usual to load the dipoles with resistance since their length can
only be about A/2 at the top end of the frequency band, thus giving a low radiation
resistance at 30 MHz.
One additional facility from a phased antenna system is that it can also provide a
substantially omnidirectional pattern for communications. The two antennas have
to be fed in phase and some matching may be necessary (broadband antenna
elements are essential for broad frequency bands such as UHF), but satisfactory
systems have been achieved for the whole UHF band on a number of aircraft on
which the number of antenna sites was limited.
VHF and UHF antennas 815
The effect of aircraft attitude changes, particularly roll, has to be considered. In
general, monopole systems on a finite ground plane are more susceptible since they
have a higher proportion of cross-polarisation.

line of
flight

\
\

port j starboard

port antenna

Fig. 17.101 Typical patterns for amplitude homing an tennas

aircraft
skin
director
sleeve dipole

Fig. 17.102 Arrangement and azimuth patterns of 200 M Hz dipole-director system

Amplitude homing: This requires the production of two mirror-imaged asym-


metric radiation patterns such as Fig. 17.101, the DF ratio being simply the ratio
of their amplitudes at a given angle.
Four commonly used methods for producing these patterns will be described:
816 VHF and UHF antennas

(i) Short Yagi: Two-element or three-element Yagis will produce the desired
cardioid pattern. They will normally be mounted near wing-tips to minimise pattern
modification by the aircraft. Dipole or monopole systems can be used, the latter
being mounted on the wing surface and being rather more sensitive to position. A
frequency range of 1.36:1 centred on 200 MHz has been achieved with three mono-
pole elements. In this instance the director is tuned for the high end of the band
and the reflector for the low end.
(ii) Dipole-director pair on side of fuselage: Fig. 17.102 shows typical patterns for
a dipole-director pair mounted parallel to the aircraft centre line and spaced about
0.2X from the fuselage side. The necessary skewed patterns depend on diffraction
around the nose and the distance from the nose has to be determined by measure-
ment on each installation. 10% bandwidths have been achieved, the limitation being
the sharpening of the DF ratio and the loss of on-course signal at high frequencies
and the converse at low frequencies.

Fig. 17.103 (a) Terminated loop antenna. (b) Azimuth pattern of vertical loop as {a)

(iii) Slot in side of fuselage: A horizontal slot in the side of a metal nose can pro-
vide a low-drag solution with good radiation-pattern shape. On a parallel-sided nose
with radome the slot should be immediately behind the forward edge of the metal,
whilst for an all-metal nose the slot central normal should be skewed 45° from the
line of flight. Cary17 gives some examples of the use of this form of antenna. For
most aircraft a lower frequency limit of about 300 MHz will apply since typical
cavity dimensions will be 0.5X long, 0.22X high, 0.1X deep for a 10% bandwidth.
Decrease of cavity dimensions brings severe restrictions in impedance bandwidth.
Slots may be probe or direct fed, off-centre feeding being used in the latter
arrangement to improve matching.
(iv) Terminated loop: A vertical loop such as Fig. 17.103 will produce in the plane
of symmetry a cardioid radiation pattern if the correct terminating impedance is
used. Typical dimensions for a full loop would be 0.03X high and 0.07X wide, but
VHF and UHF antennas 817

the antenna is more often used against a horizontal metal surface in which case the
height is halved. Experiments have shown that the shape of the antenna is not
critical, full loops in which the height and width were transposed yielding excellent
results over a frequency range of 2.5 :1.
These loops may also be used in pairs mounted on either side of a fuselage; in
this form the antenna is bisected vertically. It has been found better to use a double
loop, which might perhaps be better described as a terminated dipole, to reduce
cross-polarisation pick-up.
The half-loop under the fuselage is usually constructed to be fed at each end in
turn, the terminating resistance being switched at the same time. In this way two
mirror-imaged radiation patterns can be obtained. Some criticism of the half loop
has been based on its susceptibility to cross-polarised radiation, and it is to be
expected that this will be true on the small ground planes available in the 30-
76 MHz band where it has been mainly used. Its application to higher bands (it was,
in fact, used between 200 and 600 MHz in the Second World War) with larger loops
should be much less susceptible.

parasite

starboard

tail
Fig. 17.104 Effect of parasitic notch on VHF communications antenna patterns
100 MHz
120 MHz
140MHz

Special problems of aircraft antennas: A number of problems arise from the


form of the airframe, and among them may be mentioned:
Discontinuities
Use of non-metallic structure
Modulation from propellers and rotors

Discontinuities: Control surfaces in wings and tail necessarily imply gaps between
one metal surface and another. Such gaps, which are bridged by low-impedance
818 VHF and UHF antennas

paths at hinges, may be considered as slots if earthed at each end or as notches if


open-ended. Because the latter are in an edge of the surface, there will be a high
current density, and they can have profound effects when resonant. Fig. 17.104
shows the effect of the tail-plane/rudder gap on a fighter aircraft. The notches may
not be so obvious as this example. Fig. 17.105 shows effects of notches at the rear
of the engine fairings on VOR antenna patterns; the notches were covered with
dielectric and their existence was unsuspected. It should be noted that hinge points
are not always low-impedance junctions as dielectric bearings may be used. If the
structure exists, exploration with a grid-dip meter is advisable to search for
resonances.

[starboard

Fig. 17.105 Effect of parasitic notches on ILS localiser antenna patterns


Notches bonded
Notches unbonded

Use of non-metallic structure: The increasing use of composite materials gives


cause for concern since their electrical behaviour may be much more frequency
dependent than that of metals. Joining of panels of such materials presents prob-
lems at low frequencies and may invalidate the use of scale models for antenna
siting because of uncertainty of the behaviour of the surface of the full-scale air-
craft.
Where electrical transparencies exist — windows and glassfibre hatches, for
example — the metallic structure inside the airframe will now be strongly illumi-
nated by radiation from antennas. Electrical cables, control rods and earthing strips
may all act as efficient re-radiators in unwanted areas. It becomes a very difficult
problem to control radiation patterns under such circumstances. The use of some
form of shielding over the transparencies, either as foil, mesh or even transparent
gold film, will often need to be considered. If these are used, careful bonding to
surrounding structure is essential to avoid the formation of parasitic slots or loops.
VHF and UHF antennas 819

Propeller or rotor modulation: Two effects can be associated with moving


elements such as propeller or rotor blades. The first is variation of impedance,
which is normally only a problem when the antenna lies close to the path of the
blade. The second, changes of pattern due to re-radiation, may be more serious
particularly if amplitude-modulated radio equipment is used. There is some
experimental evidence that frequency-modulated equipment is significantly less
susceptible.
Among other problems peculiar to aircraft the following have to be considered:
Need for operation when the aircraft is on the ground
Ground clearance under fuselage
Effect of undercarriage being lowered in flight
Large control-surface movements in take-off and landing phases

17.4.7 Missiles and rockets


Whilst in many ways missiles and rockets may be considered as simplified aircraft,
the coverage requirements demand different treatment. Within the VHF and UHF
bands the main systems will be for command and telemetry, primarily the latter.
Their use will be restricted to defined test ranges on which the ground stations have
been sited to cover specific flight regions, or in the case of sounding rockets the
ground stations may be sited as required for each launch.
In the case of missile telemetry it should be possible to restrict the coverage so
that there is no requirement for signal on-axis. This is desirable since it is difficult
to achieve with a simple antenna system, and there may be a significant reduction
of signal strength due to engine efflux.
Rockets may be launched at vertical or near vertical angles, and if it is necessary
to collect data for the whole flight, the coverage requirement relative to the rocket
may be from the tail to within a small angle of the nose direction. If the rocket
spins about its axis, then near-spherical coverage is necessary.
If the vehicle is considered initially as a finite cylinder certain general comments
may be made.
(i) A monopole antenna normal to the cylinder will give some axial radiation but
will have a null along its own axis (see Section 17.4.2).
(ii) A diametrically-opposed pair of monopoles fed in antiphase will give good
broad fore-and-aft coverage but will not complement the roll-plane nulls,
(iii) Two opposed monopoles fed in phase will have an axial null,
(iv) Two orthogonal pairs of monopoles, each pair fed in antiphase and the two
pairs fed in quadrature, will add on axis and give a near-circular pattern in the roll
plane.
These observations lead to the general conclusions for a circular array that, for a
large number of elements, the gain in the direction perpendicular to the plane of
the array is only non-zero for an excitation phase progression of one cycle per
rotation round the array.
(ii) and (iv) above are both examples of this arrangement.
820 VHF and UHF antennas

If, instead of monopoles, linear or crossed slots are used, the relative signal level
in the axial and radial directions will be altered because of the different element pat-
terns. If slots are physically impracticable, heavily capacitance-loaded transmission-
line antennas (Fig. 17.19) will provide similar patterns and are very simple to
mount.
The roll-plane ripple can be reduced to any desired amount by increasing the
number of elements. Table 17.2, from Pugh, Barker and Thomas,73 gives results for
roll-plane gain and ripple for cylinders of four different diameters with arrays of
circumferential slots.

Table 17.2 Gain (dB) in roll plane


Ka S L Max. Min. Ripple, dB
5.43 8 1 + 0.7 -4.4 5.1
5.43 9 1 -0.6 -2.4 1.8
5.43 10 1 -1.1 -1.9 0.8
7.48 11 1 - 1.0 -1.9 0.9
7.8 8 1 + 0.9 -2.1 3.0
7.8 10 1 + 0.6 -3.2 3.8
7.8 11 1 -0.6 -2.3 1.7
7.8 12 1 -0.9 -2.1 1.2
7.8 15 1 -1.5 -1.5 0
7.8 12 2 + 0.3 -2.4 2.7
7.8 12 3 + 1.7 -3.8 5.5
10.02 15 1 -1.5 -1.7 0.2
Ka = circumference/wavelength
S = number of slots
L = cycles of phase progression around array

Information on the attenuation due to efflux is difficult to obtain and will, in


any case, vary according to the propellant used. Solid fuels, particularly those with
metallic particles, are regarded as giving the most attenuation.
On some small missiles there is insufficient room for flush antennas, and either
drag or handling problems prevent the use of external elements. In these instances
notches cut in the rear of wings or fins have been used. Another antenna is the nose
spike, (Fig. 17.106). As may be expected, the radiation pattern in the plane of the
missile axis is that of an asymmetric dipole and there is no possibility of obtaining
axial coverage. Provided that this is not necessary, the spike offers, on missiles less
than A/2 long, a simple and effective solution.
The general problem of coverage in missiles is considered in some detail by
Mahoney.62 Although this study was for 1.5 GHz telemetry, the general conclusions
apply to frequencies below 1 GHz.
VHF and UHF antennas 821
17.4.8 Satellites
There are two classes of VHF and UHF antennas for satellites:
(a) Telemetry and tele-command (TTC)
(b) Scientific experiments
It is not proposed here to consider the second class because of their diverse
nature, but it is worth pointing out that the aerodynamic and mechanical loads in
space are negligible and it is therefore possible to use structures, long wires for
example, which would not be mechanically practical on earth.

Fig. 17.106 Yaw-plane pattern of \ / 4 nose spike on 4 \ long slender missile

Although frequencies in the 2 GHz band are being increasingly used for TTC,
there is still considerable usage of 137 and 148 MHz, and these systems will be
described. Telemetry of the 'housekeeping' data may be required from launch, so
the antenna system requires good spherical coverage.
Widespread use is made of a four-element monopole array, loosely described as
a turnstile. It may be a true turnstile as in Fig. 17.107 or the elements may be
physically separated and set round the periphery of one face of the satellite body
(Fig. 17.108). In each case there is a progressive phase shift of 90° between adjacent
822 VHF and UHF antennas

elements. It will be recognised that this arrangement, with one-cycle phase pro-
gression, follows the scheme described in Section 17.4.7.
Much use can be made of numerical methods for determining the radiation pat-
terns. Because the satellites are normally small in terms of wavelength, magnetic-
and electric-field integral equations, MFIE and EFIE, are the most appropriate. Use
may also be made of wire grid modelling as described in Section 17.4.1. A good
summary of the application of these methods is given by Albertsen, Hansen and
Jensen.2

Fig. 17.107 Canted turnstile on satellite boom

Because the TTC antennas have to operate from launch they have to radiate
through the shroud which protects the satellite in the launch phase. In most cases
this shroud will be of glassfibre and will not, therefore, have significant effect at
VHF. Some launchers, notably uprated Thor-Delta and Ariane, have a metal shroud
with dielectric windows. Clearly radiation through these will depend on the relative
position of the satellite antennas to the apertures. The shroud can be considered as
a cavity resonator in which a complex field will be excited, particularly if the
VHF and UHF antennas 823

antennas lie around the periphery of the satellite and close to the inside of the
shroud. Determination of radiation for a given satellite is best achieved by scale
modelling.
It may be necessary to fold or retract antennas whilst the satellite is within the
shroud. The need for operating in this condition may constrain the amount of
retraction which is acceptable.

Fig. 17.108 Turnstile with separate elements

17.4.9 Antennas on persons and animals


Antennas may be required to operate in close proximity to human beings and
animals as part of a communications or a telemetry system. The use of radio track-
ing for observing the behaviour of wild animals is widespread; consideration of the
antenna is included here since there are similarities with the combination of
antenna and human body.
Personal communication equipment may be used in several different circum-
stances which affect the choice of antenna system:
(i) Military: man-pack with nominal quarter-wave whip antenna attached to the
equipment.
(ii) Public and private operators, e.g. police, fire services, mountain rescue teams,
in which the size of the radio equipment and the antenna are constrained by the
role of the operator,
(iii) Surveillance, in which the equipment and the antenna have to be concealed.
824 VHF and UHF antennas

Human body as a radiator: In the following discussion, only vertically polarised


radiation is considered, as this is used by all current systems. Bach Anderson and
Balling4 have shown that, in the frequency range 30-80 MHz, the human body
behaves in a manner similar to a lossy dielectric cylinder. Suitably fed it can act as a
moderately efficient radiator; Fig. 17.109 shows a comparison with a number of
metallic whip antennas, both man and whip being base fed above a ground plane.

-5

dB

-10

30 40 50 60 70 80
MHz

F ig. 17.109 Man as an tenna


Relative gain of base-fed man compared with base-fed whips
2 m whip
1.5 m whip
1 m whip

This concept of man as a radiating element is reinforced by measurements on a


30-70 MHz manpack antenna carried on the back (Rashid74). The antenna and
radio equipment together form an asymmetric dipole whose impedance can be
calculated approximately by the mean-value formula. Measurements show that the
impedance of the man and manpack combination is closely matched by that of an
antenna extended by a fixed amount determined by the attitude of the man, e.g.
standing or kneeling.
Bach Anderson and Hansen5 extended the theoretical study of man as a
radiator using in the first instance a single dielectric sphere to represent the body.
A subsequent analysis uses a large number of small dielectric spheres to allow for
the non-uniformity of distribution of polarisation currents in the body. Even from
the simple model of a dipole close to a dielectric sphere three important factors
emerge:
(a) the radiation resistance falls rapidly as the separation decreases once a certain
minimum distance has been reached
(b) the efficiency falls rapidly as the separation decreases
(<?) the azimuth radiation pattern becomes cardioid in form with an increasingly
deep minimum in the direction of the body.
VHF and UHF antennas 82b
Fig. 17.110 demonstrates the first point, showing the change of impedance as a
dipofe is brought closer to the body.
The azimuth radiation patterns for an antenna in front of the body show the
body acting as a director for frequencies below about 80 MHz and as a reflector for
higher frequencies. Fig. 17.111 gives typical results for frequencies between 30MHz
and 900MHz. At the higher frequencies there is a marked height dependence,
positions near the waist showing considerable reduction in all-round gain as well as
severe backward shielding. It seems probable that the coupling to the body will
vary significantly with height; results with persons of differing heights tend to bear
this out.

Fig. 17.110 Effect of body on impedance of antenna on UHF (460 MHz) portable radio
equipment

From the various studies a number of general points arise; these should be con-
sidered in designing a radio system to be carried by a man:
(i) The radio equipment has to be considered as part of the antenna system, this
may also be true for the earpiece and microphone leads unless inductive loop coup
ling is used.
(ii) If operational constraints permit, the equipment should be held as far away
from the body as possible.
826 VHF and UHF antennas

(iii) For UHF and the higher VHF bands the advantage to be gained by having the
equipment as high up the body as possible is greater than simple height gain would
predict.
(iv) When the antenna has to be in close proximity to the body, the impedance will
vary rapidly with position. The basic antenna should have as broad a bandwidth as
possible. Short, fat, helical antennas radiating in the side-fire mode have been
shown by several writers to be most effective. Simple tuning means should be
provided unless the loss due to mismatch can be tolerated. Providing adequate
tuning requires a knowledge of the antenna impedance. Measurement requires con-
siderable care to ensure that the measuring cable does not become part of the
antenna, Krupka55 and King and Wong48 show similar choke systems to minimise
this effect, although Kuecken,57 discussing HF measurement, states that chokes
were ineffective, probably because of the very high impedances involved. In any
case the system must be designed 'on the body' and not in isolation.

front

left right

Fig. 17.111 Azimuth patterns of a vertical antenna on the front of a standing man
82MHz 155MHz
245 MHz 800 MHz
Scales are relative only. No relation between different frequencies

Antennas for radio tracking: For ease of installation antennas fitted to mammals
usually consist of a loop fitted round the neck and carrying the transmitter and bat-
teries. These are arranged to lie at the lowest point so that the antenna is fed as a
vertical loop. Evidence from field trials is that on the animal the antenna is
VHF and UHF antennas 827

essentially vertically polarized in the horizontal plane and has a directional response
as a free-space loop. This indicates that coupling to the animal is small. This type of
antenna has been used with transmitters in the 27, 101 and 170 MHz frequency
bands, the higher frequencies normally with the smaller animals.

17.4.10 Methods of measurement on vehicles


When the performance of the antenna is significantly affected by the vehicle on
which it is mounted, measurement work in the design or performance proving
stages must include as much of the vehicle as is relevant to the particular measure-
ment. This can often add considerably to the cost and difficulty of making the
measurement; this Section examines the most economical methods for typical
installations.

Radiation patterns: Wherever possible use should be made of scale models, but,
particularly with land-based vehicles, the complexity of the model may make it
cheaper to buy, for example, a used motor car which will also be valuable for
impedance measurements.
Measurements on land-based vehicles can be most rapidly done by rotating the
vehicle on a flat-bed turntable. Even the type of turntable used in car showrooms
will, if equipped with a position synchro, give good service out of doors for years.
In setting up the antenna test site the normal criteria for ground ranges (see
Chapter 8) should be observed. The antenna aperture D is to be taken as the
largest dimension of the vehicle. Some reduction in range below the far field
distance 2D2/X can be permitted at the expense of definition of depths of nulls.
This may be unimportant for most communications antennas.
In the absence of a turntable the technique is to keep the test vehicle fixed
and to take a measurement antenna round it at a fixed distance. Ideally, a con-
tinuous plot of signal level around the cricle is required. This can be done if the
measurement antenna is fixed to a vehicle driven round the circle, means being
provided to convert the vehicular motion to angular position with respect to the
test vehicle. A simpler system, adequate for most purposes in this frequency range,
is to define successive 10° points round the circle and to measure at each in turn.
If in addition significant maxima or minima are noted between the 10° points, a
good representation of the radiation pattern can be obtained. With practice it is
unnecessary to halt at each 10° marker and the vehicle can be driven continuously
round the circle. It should be possible to measure one pattern in 10 minutes, very
much more quickly than trying to turn the test vehicle to successive headings.
The same technique may be used for aircraft provided that the antenna is not
mounted on a lower surface of the aircraft. As an indication of the extent to which
the range can be reduced, Fig. 17.112 compares a 100ft aircraft measured at a
radius of 200 ft centred on the antenna, in the aircraft nose, with the far-field
scale-model pattern.
Flight polar diagrams on aircraft should be used only when scale-model or
ground measurements are impractical. Patterns can be measured either by flying
828 VHF and UHF antennas

an orbital or a daisy-pattern flight path. In each case the object is to present diffe-
rent attitudes of the aircraft to a ground station at which the field strength is
recorded. Fig. 17.113 illustrates the two techniques. The daisy-pattern usually con-
sists of 36 runs at 10° intervals over the pinpoint; this is time-consuming but has
the advantage that the aircraft is flying straight and level over the pinpoint and the
measuring range remains constant. In the orbit the object is to fly as flat a turn as
possible; this may mean a radius of several kilometres and hence the range has to be
increased to minimise the correction for the radius of the aircraft turning circle.

nose

port starboard

tail

Fig. 17.112 Comparison of full-scale and model radiation pattern measurements


Azimuth radiation pattern of ILS localiser aerial in nose radome
Full-scale frequency: 112 MHz
Full size aircraft measured on ground
1/15 scale model

The ground station will normally be elevated to give adequate range so that the
graph of signal strength against range is of the form shown in Fig. 17.114. The aim
should be to choose a pinpoint at P so that range errors have the least effect on
signal strength. For a detailed treatment, see Bothe.8

Impedance and VSWR: Vehicles rarely consist of isolated flat sheets of metal,
so measurements on flat ground planes are not representative of antenna perfor-
mance on a vehicle. It will usually be necessary to use either the vehicle or a
mock-up of it for impedance measurement. To what extent need the whole vehicle
be simulated? Consider a monopole mounted at the extremity of an aircraft tail fin
- clearly the antenna and the remainder of the aircraft constitute an asymmetric
VHF and UHF antennas 829
dipole. From Section 17.2, the impedance of such a dipole is approximately the
sum of the impedances of its two components, one of which is the antenna itself
and the other is the remainder of the aircraft. Non-circular sections can be trans-
formed to equivalent circular cross-sections and the resulting model of the aircraft
is of a series of cylinders and cones of varying diameter. It will be seen that, for
most frequencies, the diameter exceeds 0.05X and reference to measured
impedances of such fat dipoles, for example Brown and Woodward,10 shows that,
for arms greater than A/2, the impedance of the dipole is practically invariant. It
therefore follows that, provided the length of the structure exceeds A/2 from the
base of the antenna, the measured impedance will be a good approximation to
that on the whole aircraft.

ground
station
start

ground
station

Fig. 17.113 Techniques for flight polar-diagram measurement


Du D2: chosen in accordance with Fig. 1 7.114
(a) 24-point daisy pattern
ib) Orbit. R <0.03O 2

Similar arguments can be applied to other continuous structures. For an aircraft


fuselage, sufficiently fat, the extension either side of the antenna need not exceed
A/2. The argument cannot be applied to an antenna mounted normal to an aerofoil
section, the performance here resembling that on a flat sheet and hence being much
more dependent on sheet size.
Apertures in vehicles may permit coupling into the interior with an appropriate
modification of the impedance. As the size of the internal cavity may permit multi-
moding, it will not normally be possible to calculate the effect, so any mock-up
should include the apertures and a reasonable representation of the interior.
830 VHF and UHF antennas

Gain: Gain for vehicular antennas is usually referred to some agreed standard.
For a motor-car it might be a X/4 monopole at the centre of the roof, a halfwave
dipole or a ground-plane monopole. It may not be important to know absolute gain.

-80

-90
useful range
£ -100 of P

~° -110

I -120
a

•o -130

-150 20 40 60 80
range km

Fig. 17.114 Typical range-power curve for constant antenna heights and frequency

Since the measurements have to be made on ground reflection ranges it is


important that the reference antennas have an elevation pattern similar to that of
the antenna under test. The measuring point must be at the peak of the field pat-
tern. It is usual to make substitution measurements at a fixed distance; if it is
possible to make measurements at a number of distances, two asymptotic curves of
signal level will be obtained and some of the uncertainty due to ground reflections
can be removed.
It should be obvious that conductors extending from a vehicle may modify its
antenna performance. Wherever possible, therefore, internal measurements should
be made using equipment with self-contained power supplies. This is particularly
important at the lower end of the VHF band.

17.5 Position-finding antennas

This Section describes antenna systems used either to determine the direction of
an incoming signal or to lay down defined patterns along which a vehicle,
particularly an aircraft, can be steered.

17.5.1 Homing systems


The general principles of homing systems have been described in Section 17.4.6.
There the need for equality of impedance and symmetry of the antenna system
was stressed. It is only intended here to show the dependence of monopole sys-
tems on the shape of their ground plane, from which the superiority of balanced
VHF and UHF antennas 831
dipole systems can be deduced. Fig. 17.115 compares the phase shift, as a function
of angle, of two identical antennas mounted on a circular ground plane and on the
roof of a road vehicle. In both instances the antennas were mounted symmetrically.
Problems of asymmetrical scatterers are considered in Section 17.8.

120
-o TOO
j? 80
$ 60
o
.c
a
20
-180 azimuth bearing
0 90 • 180
20

Fig. 17.115 Measured phase shift between pair of antennas showing influence of mounting
surface
Isolated ground plane
Vehicle roof of same size

17.5.2 Tracking and direc tion-finding sy stems


The function of a tracking or direction-finding system is to determine the direction
of a target emitting radio-frequency signals and moving relative to the position of
the tracking station. The tracking required may be to follow a wild animal equipped
with a radio transmitter, or a satellite or a missile telemetry and control system, so
that a wide range of accuracy and angular movement may be involved. Strictly,
perhaps, direction finding defines the elevation plane containing the target, whilst
tracking defines a line in a plane, but in this frequency range the difference is not
significant.
In the simplest case of animal tracking great accuracy of bearing is not required
and would in any case be impracticable because of the terrain. It is therefore the
practice to use some form of directional antenna, a Yagi for example, and to orient
it for maximum signal. For convenience at frequencies below 100 MHz a dipole pair
or H antenna is often used, the elements being loaded to reduce their size. There is
no fundamental reason why active elements should not be used provided that
variations in gain can be kept to a minimum.

Elevated H-Adcock array: The next system in terms of simplicity is the elevated
H-Adcock array, Fig. 17.116. It consists of a pair of balanced dipoles coupled
832 VHF and UHF antennas

together to produce a pattern with two nulls, plus a centre element whose output
can be appropriately combined with the pair to produce a cardioid pattern to
resolve the ambiguity. The array is raised above ground as high as possible, the limit
being determined by the need to avoid a null in the elevation plane in the angular
range of interest. This null would arise from the combination of the antenna and its
image in the ground. If h is the height of the antenna system above ground and 0
the angle of elevation, the first null occurs at sin 0 = X/4/z.

D D

Fig. 17.116 Arrangement of elements for H-Adcock array (rotating)


Outer elements combined in phase

The pick-up for the Adcock pair is given by the expression


EA = 2Ehe sin (Kd sin d) cos u>t
where E cos cot is the electric field at the centre of the antenna, he is the effective
height of each dipole, d the half width of the system, and 6 the angle of the wave-
front with the line joining the pair.
Fig. 17.117 gives values for this expression and shows that useful patterns can be
obtained for 2d/A> 0.75. It is advantageous to use the largest possible spacing to
minimise mutual coupling between the two dipoles and the third antenna and also
the support mast. It will also be noted that the larger spacing provides a sharper
null.
The whole antenna is rotated to find the null directions. The ambiguity is
resolved by adding to the output an equal amplitude from the omnidirectional
sense antenna. Since the directional output lags 90° on the incident field whereas
the omnidirectional output is in phase, a 90° phase shift has to be added to one
VHF and UHF antennas 833
or other component so that the two components are in exact phase or antiphase.
The resultant is a cardioid having a maximum at 90° to the minimum of the direc-
tional signal; by noting whether the signal increases or decreases when the sense
antenna is switched in, the ambiguity can be resolved.
If the amplitudes and phases of the directional pair and the sense antenna are
not equal, the front/back ratio of the cardioid becomes finite and may becomes so
small that a good sense indication is impossible. Griffith and Rosinski29 have dis-
cussed the factors controlling the performance of a VHF system in more detail
than can be given here.

signal level , V

Fig. 17.117 Equatorial-plane patterns of pair of equal, co-phased dipoles as function of


spacing

20= 1.0\

Fixed Adcock system: Instead of rotating a directional pair it is possible to


obtain bearings with two fixed orthogonal pairs as shown in Fig. 17.118. By
comparing phase and amplitude of two outputs as well as that of the centre sense
antenna, the bearing of the transmitter can be determined.
In this, as in the rotating system, equality of impedance is very important.
Broad-band, and therefore fat, dipoles are preferred for this reason.
834 VHF and UHF antennas

The small base line of both forms of Adcock means that a good reflection-free
site is required to eliminate bearing errors since all the antennas are equally
affected. Earp and Cooper-Jones25 pioneered the use of the commutated antenna
system in what is now known as CADF (commutated antenna direction-finder) or
DDF (Doppler direction-finder) equipment.

Fig. 17.118 Fixed Adcock array

Commutated antenna system: Fig. 17.119 shows a typical arrangement of 18


dipoles equally spaced on a 2X diameter circle, the antenna-to-antenna spacing being
approximately one-third of a wavelength. Sampling signals around the circum-
ference at constant angular velocity gives a Doppler shift varying sinusoidally with
angular position and a phase shift equal to the azimuth. It is then possible to
resolve the bearing by normal phase-measurement techniques. A reduction in RMS
site error by a factor of 7 compared with an Adcock system has been achieved with
the system described above.
The system shown in Fig. 17.119 uses dipoles. Monopole systems have been
developed for use in the airborne UHF band 225-400 MHz. Although the ground
plane helps to reduce high-angle reflections, there is a small error in phase shift
since each antenna element is asymmetrically disposed on the ground plane and
has therefore a phase-centre variation with azimuth.

Tracking in two planes: The antennas described above are used essentially for
tracking in one plane. The need for VHF and UHF tracking in two planes originated
with the radio astronomers and has its major application with satellites and ballistic
missiles. Two separate approaches have to be considered:
(i) Orthogonal arrays of omnidirectional elements whose phase is varied continu-
ously to sweep the beam.
(ii) Circular arrays of comparatively narrow-beam antennas combined to give a
conical scan.
The basis of the first approach is the interferometer.
VHF and UHF antennas 835
The interferometer: Consider a pair of simple antennas spaced S wavelengths
apart and fed in phase, Fig. 17.120. Their radiation pattern in the plane through the
pair consists of a number of narrow lobes which are the product of the individual
element patterns and the grating pattern given by two spaced sources. Around the
central lobe the width between the first nulls is 513 S degrees. It will be seen that
the beam width of this array is half that of a continuous array of aperture S. By
judicious choice of the element patterns and 5, typically 5 = 3 to 5 times the aper-
ture of the individual antennas, it will often be possible to obtain adequate angular
accuracy by seeking the peak value of the largest lobe without the need for any
phasing. This implies that the whole assembly can be rotated or that the move-
ment of the target is such that the interferometer can be set so that the target
passes through successive beams.

reference
antenna

Fig. 17.119 Plan view of 1 ^-element commutated DF antenna


Elements are \ / 2 dipoles

By altering the phase between the two elements of a pair, the patterns can be
swept in the plane of the array as described by Ryle.76 Kraus54 describes a simple
form as a phase-switched interferometer. For the arrangement of Fig. 17.120, when
the antenna elements are in phase, the far field pattern is
836 VHF and UHF antennas

where E0((j)) is the normalised pattern of an individual element and \// = 2nS sin 0.
With the phase to one element reversed the pattern becomes

which represents a lobe structure having a null at <p = 0 instead of a maximum.

element pattern

interferometer
pattern

SA

Fig. 17.120 Equal-element interferometer

Since a null position can be determined with more accuracy than a peak, the
phased-switched system would be used in the following way. The source direction
would be determined coarsely from the position of the greatest peak of the in-phase
system. With the system switched to anti-phase, the source position now lies within
the central null and can be located more accurately.
Fig. 17.121 shows an interferometer in which the two elements are dissimilar.
A special case of this arrangement (Fig. 17.122) in which S = 0 is known as the
Mills Cross (Mills and Little67). It forms the basis of a number of radio-telescope
antennas, some of which are described by Kraus.54
By using other pairs of elements at different spacing, the side-lobe structure of
an interferometer can be suppressed so that the position of a distant source can be
determined with an accuracy only limited by the size of the baseline and the fre-
quency stability of the system.
An alternative use of the system, employed by a number of satellite ground
stations, is to use a close-spaced pair of antennas to give a coarse angular position
and a much wider-spaced pair to define position within the wider beam. Clearly
any convenient form of antenna element can be used. Among those described by
VHF and UHF antennas 837

Kraus are helices and dipoles, with and without parabolic reflectors. For satellite
tracking systems where the required accuracy permits rather simpler systems, slotted
cylinders have been used. These, based on the work of Alford, provide an aperture
of 2X from a single feed. The decision to use a few complex elements or many
simple ones must depend, in part, on the reliability and consistency of phase of the
feed system under all weather conditions.

E/rf)

-H

Fig. 17.121 Asymmetric interferometer

Fig. 17.122 Mills cross


Note: All elements similarly polarised

17.53 Transm it ting an tennas for navigatio n system s


Two important navigation systems for aircraft operate in the VHF-UHF range.
These are ILS (instrument landing systems) and VOR (VHF omnidirectional
range). The first is an approach and landing aid in which equisignal paths are
838 VHF and UHF antennas

provided at defined azimuth and elevation angles;the second is an en-route navigation


system which permits the aircraft to determine the bearing of the ground station.
Whilst the ILS antenna requirements are similar to those of a homing system, those
of the VOR system are analogous to a direction finder.
ILS has three components:
Localiser (Azimuth guidance) 108-112 MHz
Glidepath or glideslope (Elevation guidance) 328-336 MHz
Marker (En route position) 75 MHz
whilst VOR operates in the range 108-118 MHz.
Both systems use horizontal polarisation.
Since ILS uses conventional antenna elements — dipoles and loops with or with-
out reflectors — it is not proposed to describe them further. The significant features
of ILS are in the methods of applying the two modulating frequencies to each
antenna element. A typical system is described by Garfield and Norbury.27
The VOR antenna system has to radiate two signals modulated at 30 Hz. One,
the reference signal, has constant phase and amplitude whilst the phase of the other
varies with the direction of propagation. Two forms of VOR are in general use,
distinguished as 'conventional' and 'Doppler' VOR.

Conventional VOR: A typical conventional type has been described by Popp72


and its significant features are summarised below. The omnidirectional antenna
consists of four short dipoles arranged in a circle. Each dipole, which includes some
inductive loading, is balun fed, the four outputs being connected to a coaxial cable
at their common point. The azimuth pattern deviates from circular by less than
± 0.25 dB.
The rotating pattern is provided by crossed dipoles, 0.17 A long, each connected
to one output of an electronic goniometer. The dipoles are sleeved and with some
capacitance end-loading. They and the 4-dipole loop antenna are printed on a single
copper-clad dielectric board for consistency.
By itself this system would also have a vertically-polarised component increasing
with elevation angle. The bearing information of this component is 90° out of
phase with the horizontal component, and so would give rise to bearing errors
dependent on the aircraft-antenna radiation patterns. It is difficult to eliminate a
vertically-polarised component in the aircraft antenna, so the ground antenna has a
cage of 24 vertical rods on a 1 m diameter surrounding the antenna elements. The
cage is 1.8 m high and is surmounted by a cage 0.9 m high with 12 rods. This mini-
mises radiation from the top plate of the main cage.
To control the elevation pattern the antenna system is mounted on a counter-
poise whose diameter and height above ground are important. Either a single
element or stack of two may be used, depending on gain and elevation requirements.
Fig. 17.123 shows the main dimensions of a 2-element system.

Doppler VOR: Like the Doppler direction finder, the direction-dependent


VHF and UHF antennas 839
signal is generated by moving a radiation source on a circle. Instead of mechanical
rotation, a number of antennas around the circumference of the circle are sequen-
tially fed. To give the appropriate frequency deviation of ±480 Hz in a single
rotation of 1/30 second, the diameter of the circle is approximately 5A, or 13.5 m.
This permits an improvement in course stability of the order of 10 compared with
a conventional VOR, and thus allows the system to be used in sites which are too
poor for the conventional system.

18 rod 3-6
polarisation
cage
[cut away to show feeds]
2-7

1-8
feed elements in
radomes

-I
0-9
5 0 dia

ground plane equipment


room

25

/ / / / / / / / / / / / / / / / / / / / / / / / / / /

Fig. 17.123 Two-stack VOR transmitting antenna


Dimensions in metres

A typical system described by Crone and Popp,21 uses 39 elements in a 13.5 m-


diameter ring with a similar, central element as reference antenna. All are mounted
above a 40m-diameter counterpoise. It would have been possible to use the
4-dipole loop described above, but since phase stability is a more severe constraint
in this form of VOR, a rather larger Alford loop is used. This gives a variation of
± 0.5 dB in amplitude in the azimuth plane and ± 5° in phase.
840 VHF and UHF antennas

17.6 Dielectric-loaded antennas

Two different aspects of dielectric loading are considered here. The first deals with
attempts to reduce the physical size of antennas by loading and the second with the
use of dielectric materials as protection, where the aim is to minimise the influence
of the dielectric.

17.6.1 Dielectric loading


The search for a loading material which would permit a dramatic reduction in the
size of antennas has continued since antennas were first used. While most work has
been with low-loss dielectric materials, there has recently been a recognition that
magnetic material may be equally useful.
In a dielectric filled coaxial transmission line the electrical length is increased by
a factor \fe and it is reasonable to expect that a similar increase would occur with a
dipole immersed in a dielectric. This is true only if the dielectric is infinite in
extent, any other arrangement constituting a stratified dielectric sheath in which
the outer layer is air. The increase in electrical length for a given physical length will
thus be less than y/e.
M. S. Smith85 identifies three distinct regions for dielectric loading. These are:
(i) Electrically small antennas in which the main effect is capacitive, and the coat-
ing may be thick or thin.
(ii) Near resonant antennas with thin coatings, in which the main effect is a
moderate reduction in resonant frequency and input impedance,
(iii) Thick dielectric coatings; a dielectric resonance effect has been identified, giving
narrow bandwidths and low radiation resistances.
As an example of class (ii) Smith quotes a monopole 100 mm long. Without
coating, the resonant frequency was 750 MHz; with e = 40 and overall diameter of
10mm the resonant frequency was 450 MHz, but with a 3 dB loss in efficiency
assuming lossless matching to 5012.
Some measurements with thick coatings and short elements [class (i), James and
Burrows35] show that, while it is possible to resonate a monopole as short as
0.05X, the radiation resistance is of the same order as the uncoated monopole.
Whilst the bandwidth (defined as the frequency range over which R > ± X) is of the
order of 15%, the low impedance (R — 512) means that the instantaneous bandwidth
on a 5012 system is narrow and significant matching losses must be considered.
Later work by James and Henderson36*37 suggests that, when the electrical
length in the dielectric is an exact number of quarter wavelengths, resonator action
occurs in the dielectric. Depending on the relative diameter of monopole and
dielectric (Fig. 17.124), there will be some end-loading which will increase the
effective length of the antenna. Their calculations indicate that greater height
reduction can be obtained with material of high permeability than with material
having numerically equal permittivity and permeability, Fig. 17.125. Examination
of impedance curves also shows a higher resistance at resonance with ferrite material
VHF and UHF antennas 841
than with high dielectrics. Loss factors have been computed for various materials;
these show (Fig. 17.126) that, for a given loss factor, the greatest height reduction
is achieved when fxr = e r . Some caution must be used in interpretation of this chart
since LA = 10 log10 (RA/RA + RL) where RA = radiation resistance andRL = loss
resistance. Further losses will be involved in matching these impedances to
transmission-line impedances, and this type of antenna would be well suited to
applications in which no transmission line is required, e.g. handheld radio sets or
integrated antenna/receiver systems.

dielectric

monopole

Fig. 17.124 Monopole in dielectric cylinder

Table 173 Performance of loaded rectangular cavity slot antennas


Loading Air Ferrite powder Solid ferrite Dielectric
Size (in) 3 0 x 7 | x 10 1 2 x 3 x4 5x2x1^ 12x3x5
Volume (in3) 2250 144 15 180
Bandwidth at 22 MHz 19 MHz 10 MHz
VSWR = 3
Efficiency 90% 65% 30% 85%
Weight (lb) 2Si 16| 3.6 14.5
Frequency (MHz) 300 320 352 316

The results obtained above refer to coated monopoles on a ground plane. Yagi
and log-periodic antennas could be similarly treated to give a reduction in element
length even if reducing the axial length of the antenna was impractical for most
842 VHF and UHF antennas

frequencies. Work has also been done on reducing the size of cavity slot antennas
by loading with ferrite or dielectric. Table 17.3 due to Lyon and Hiatt60 gives
typical results.

0.03

0.02

0.01

W
1 5 9
N

Fig. 17.125 Height reduction factor N for various dielectric coatings (computed)
A: M r = 8 1 e r = 1
B: M r = 9 er = 9
C: ju r = 1 €r= 81
b/a= 20, see Fig. 17.124
tan 6e = tan SM = 0

17.6.2 Dielectric covers


A dielectric cover over an antenna may be necessary for any of the following
reasons:
(a) to prevent ingress of rain, snow, dust, sand or industrial contaminants
(b) to reduce wind resistance
(c) to provide a pressurised container around the antenna to minimise voltage
breakdown at low pressure
(d) to prevent damage to the antenna
(e) to prevent damage to personnel.
VHF and UHF antennas 843

This Section is only concerned with covers whose electrical thickness is small,
and not with radome design. Although, the covers are 'thin', they can nevertheless
influence the gain and impedance of the antennas they surround. The effect is
greatest when the dielectric is close to the high-impedance points of the antenna.
Thus a dielectric cover in the plane of a slot antenna will have a significant effect
not only in lowering the working frequency but also in decreasing efficiency if the
material is chosen for its structural rather than its electrical properties.

lOr

dB

height reduction N

Fig. 17.126 Loss factor [computed) for various dielectric coatings


A: M r = 8 1 , e r = 1 tan 8^ = 0.002
B: Mr= 9, e r = 9 tan 6e = 0
0:^=1, er=81

tan = 0

To minimise these effects the cover has to be moved to a point of lower voltage
gradient. Fig. 17.127 shows some arrangements which will make the thickness and
position of the cover less critical. It must be remembered that rain, ice or snow will
effectively increase the thickness of the cover, so the spacing has to be chosen to
allow not only for manufacturing and positional tolerances but for the anticipated
build-up of moisture in one form or another, including condensation. If the antenna
design is very critical to mismatch, it may be necessary to provide hot air to keep
the weather loading within acceptable limits.
It should be obvious from the previous Section that the effect of dielectric
covers on electrically short antennas is more significant than on self-resonant ones.
844 VHF and UHF antennas

It is true that the cover may reduce the VSWR; this is probably because of its high
loss resistance. The notch antenna in a flat sheet or sharply curving surface is
another type significantly affected by the proximity of a cover.

structural dielectric

low loss rigid foam

Fig. 17.127 Methods of reducing effect of dielectric cover


(a) Flush cover for aircraft
(b) Cover for ground installation
(c) With foam packing
VHF and UHF antennas 845

17.7 Circularly-polarised antennas

Circularly-polarised antennas may be required in a number of circumstances for two


different reasons:
(a) To minimise fading by rejection of ground-reflected signals
(b) To improve coverage
An example of (a) is ground-to-aircraft communication via satellite. When the
satellite is at a low elevation angle there is strong reflection from the ground, more
particularly the sea, in addition to the direct path. The low gain, and hence broad
beam, of the aircraft antenna does not provide sufficient discrimination by itself
and severe fading can occur. If a circularly-polarised antenna system is used the
ground-reflected signal is polarised in the opposite hand, and this provides a
significant improvement in signal quality.
The second class includes aerospace vehicles where a manoeuvre of the vehicle
alters the polarisation of its antennas relative to that of a fixed ground station.
The use of circularly-polarised ground antennas helps to minimise loss of signal
which would occur at high angles of bank as well as that due to ground reflections.
Purely linearly-polarised antennas have fundamental nulls, such as that along the
axis of a dipole, so a case can also be made for circularly-polarised antennas on the
vehicle.

17.7.1 Definition
The plane of polarisation of a radio wave is defined by the direction of the electric
vector during at least one complete cycle. In general, both the magnitude and
direction will vary during the cycle. In a plane perpendicular to the direction of
propagation, the electric vector will describe an ellipse and the wave is said to be
elliptically polarised. Two special cases of elliptical polarisation are of particular
importance:
(a) Linear polarisation in which the electric vector always lies along a fixed line;
ratio of major to minor axis of ellipse is infinite
(b) Circular polarisation in which the electric vector describes a circle; ratio of
major to minor axis is unity
The ratio of the axes (axial ratio) is known as the ellipticity, usually expressed in
decibels, and is the parameter used to define the quality of a nominally circularly-
polarised antenna.
The electric vector can rotate in either direction. The convention adopted by the
IEEE is that, viewed in the direction of propagation, the polarisation is right-hand
for clockwise rotation, left-hand for counter-clockwise (see Chapter 1). Fig. 17.128
shows the orientation of the electric vector during a complete cycle for right-hand
circular (RHC) polarisation and gives a clue to methods of propagating such a wave.
Clearly two orthogonally polarised antennas fed with equal amplitude and having a
phase shift of n/2 will generate the appropriate field. The antennas may lie in the
846 VHF and UHF antennas

same plane orthogonal to the direction of propagation and be fed with a TT/2 shift
or they may be displaced A/4 along the direction of propagation and fed in phase.
The following Section includes many of the current methods of producing circular
polarisation.

0 rr/2 TT 3rr/2 2TT

Fig. 17.128 Orien tation of electric vector for righ t-hand circular polarisation

I /0°

Fig. 17.129 Dipole and loop combination for circular polarisation

17.7.2 Combination of elements


Whilst in theory circularly-polarised radiation can be produced over a complete
sphere, the complete isotropic antenna cannot be realised in practice. Two classes
of circularly polarised antennas can be distinguished:
(a) Circularly polarised in a plane orthogonal to the antenna axis
(b) Circularly polarised axially

Orthogonal plane circularly polarised: The simplest example is the combination


of dipole and loop (Fig. 17.129). Provided that the loop is less than 0.6A in
diameter and the dipole is less than X/2, circular polarisation will result if both
elements are fed in phase and with equal amplitude. The fields of the two elements
VHF and UHF antennas 847
are then automatically in time-phase quadrature. Since both elements have the same
radiation patterns the field is circularly polarised at all points, but is zero on the
dipole/loop axis. Equal amplitude implies equal impedance, and the very different
impedance characteristics of a loop and a dipole make it difficult to provide wide
band width by this method. Practical forms of this combination are often as Fig.
17.130 due to Kandoian.43 This is a combination of an Alford loop and a mono-
pole. Since both the loop and the monopole are coaxiaUy fed, this is physically a
very convenient form of antenna. By proper choice of the impedance of the lines
inside the Alford loop, a convenient impedance can be obtained to match that of
the dipole, Bandwidths of 10% have been claimed for 2:1 VSWR, but the axial
ratio is not quoted.
Whilst it should be possible to combine a vertical dipole and slotted cylinder by
making the dipole elements large enough to act as the cylinders, this must result in
an antenna one wavelength high with some difficult matching problems. It would
be essentially a narrow-band device.

F ig. 17.130 Monop ole and AIford loop

TT/2

TT
3TT/2

Fig. 17.131 Two arrangements of crossed dipoles with relative phases

Circularly-polarised axially: Perhaps the simplest form of such antennas con-


sists of a pair of crossed dipoles with time or spatial phase shift (Fig. 17.131). When
the dipoles lie in one plane the arrangement is known as a turnstile antenna (see
848 VHF and UHF antennas

Section 17.2.1). It should be noted that in one axial direction the turnstile is
polarised RHC but in the opposite direction it is LHC. Like most of this class of
antenna the polarisation degrades to linear in the orthogonal plane. Some improve-
ment in angular coverage for a reasonable axial ratio can be obtained by using vee
dipoles instead of straight ones. A whole family of antennas based on the dipole can
be used including crossed Yagis and log periodics. For good axial ratios it is essential
that the relative phase and amplitude of the orthogonal elements should remain
constant over the desired bandwidth, which implies careful control of element
design and tolerances. It is also necessary to maintain the quadrature phase shift
between the orthogonal components.

cavity
wall p p

Cross-section at A-A showing


ridge in line of probes P

Fig. 17.132 Crossed slo ts with ridged cavity


Ht: 180° hybrids
H 2 : 90° hybrid

Various arrangements of slots with backing cavities can be used to provide


circular polarisation from low-profile antennas. The simplest arrangement is the
crossed slots of Fig. 17.132 (King and Wong47). Various feed systems can be used,
one of the neatest being coaxial cables tapped across each slot at the 5012 point,
the two cables then being connected in quadrature. An alternative arrangement
with a ring of slots is shown in Fig. 17.133. The four folded dumb-bell slots are of
reduced length, so that the spacing between parallel slots can be kept as small as
possible to broaden the beamwidth. The four slots are fed in 90° phase progression.
This arrangement gives 8 dBi gain on axis and a bandwidth of approximately 7%.
The 4 dB beamwidth is 65°. Outside the main beam the ellipticity degrades rapidly
and the reverse hand of polarisation occurs, but the antenna has application within
its main beam.
VHF and UHF antennas 849
Since a slot and a parallel dipole have orthogonal polarisations but identical radi-
ation patterns it is tempting to combine these to give circular polarisation and
various arrangements have been tried. The simplest is a dipole X/4 above the slot,

I///////.// ///7/\

'o-

270° 90°— 0-43A

180°
I
1 ft 1
[Z / / / / /

n
*/ / / / / / / A

0-53A

Fig. 17.133 Array of four folded dumb-bell slots

Fig. 17.134 Slot with parasitic monopoles

both elements being fed (Cox and Rupp20). A variation on this (Clavin etal.1*) has
a slot with parasitically coupled inverted L monopoles on either side (Fig. 17.134).
Other arrangements include a dipole skewed slightly from the slot axis to give
appropriate coupling. Physically the neatest arrangement is that due to Sidford80 in
850 VHF and UHF antennas

which the dipole lies in the plane of the slot. As Fig. 17.135 indicates, the dipole is
a folded type coaxially fed and coupled to the slot by one or two capacitive tabs.
The whole slot/dipole assembly can be printed on one double-sided board. With a
backing cavity no more than A/12 deep an impedance bandwidth of 10% (to 2:1
VSWR) is possible. This antenna, which has been named the Sidpole after its
inventor, has a good axial ratio over a very wide E-plane, but the coupling between
the slot and the dipole has to be adjusted experimentally and is critical.

coaxial feed orthogonal


to plane of slot
folded dipole
in plane of slot

microstrip feeder coupling capacitor

Fig. 17.135 Prin ted slo t-dipole {dipole-fed)

17.8 Multiple installations

The need to mount a large number of antennas in close proximity occurs in a


surprisingly large number of circumstances. To take a domestic example, one
may consider the forest of television and radio receiving antennas on the roof
of a multistorey block of apartments. This is an instance of uncontrolled instal-
lations, and it is reasonable to suppose that the performance of each antenna
system will be degraded by the proximity of the others.

17.8.1 Effect of one antenna on another


The effect of one antenna on another has to be considered in terms of three factors:
Impedance
Radiation pattern
Coupling
Figs. 17.136 and 17.137 give mutual impedance between parallel and collinear
halfwave dipoles of thin cross-section. Theoretical data are not available for 'fat'
VHF and UHF antennas 851
elements and must be experimentally determined. Calculations for parallel antennas
of unequal size appear in the literature but are of rather limited value because they
refer to thin antennas which are rarely encountered in this frequency range.

80 A/2

<n - 4 0 —
£ 0 0-2 0 4 06 0-8 10 1-2 1 4 -16 18 20

5/A

Fig. 17.136 Mutual impedance between two thin parallel \/2 dipoles

2 10

-10
0-2 0 - 4 0 6 0-8 1 0 1-2 1-A 1 6 1-8 2 0

Fig. 17.137 Mutual impedance between two thin coilinear \/2 dipoles

Walkinshaw88 gives values for the impedance of a thin dipole in the presence of
parasitic elements; a study of his figures will show that the parasitic element has
to be within A/2 to produce a significant effect on the impedance of the driven
element. It should be noted that the parasitic dipoles have a zero centre-point
852 VHF and UHF antennas

impedance. In a situation such as Fig. 17.138, where 1 and 2 are dipoles resonant at
different frequencies/! and/ 2 , the impedance of the radio equipment ZS2 a t / t
referred to the centre of dipole 2 may make this dipole a very effective parasite to
dipole 1, and vice versa. It cannot be assumed that because ZS2 is matched to the
transmission line at f2 the same will be true at f\ and in general this is most unlikely.
The effect of dipole 2 on dipole 1 cannot therefore be determined without a
knowledge of the impedance presented to its terminals at the relevant frequency.

Li 1 11 12

Fig. 17.138 Genera/ arrangement for coupling between two antennas operating at different
frequencies

Walkinshaw's paper also gives some indication of the effect of one dipole on the
radiation pattern of another. It is clear that the maximum effect occurs when the
elements are parallel and will be very much reduced if they are collinear or ortho-
gonal. When one antenna is skew to another, although the mutual coupling will be
reduced, there will be some re-radiation from the skew element in cross-polarisation,
the significance of which will depend on the system operation. In homing and
direction finding systems the presence of another antenna may assume much
greater importance. If the antenna is on the axis of symmetry of a two-element
homing system then the effect on each element will be the same. The angular
change of phase between signals received by the two homing antennas may be
disturbed, but the equiphase line will remain on the axis of symmetry. If the third
antenna is offset, however, the responses of the two homing antennas will not be
mirror-imaged about the axis and a 'squint' will develop. In the previous paragraph
the effect of terminal impedance on re-radiation was noted. This indicates that it
may be possible to choose the base impedance of the third antenna to minimise its
effect within the operating frequency range of the homing system.
The coupling between two antennas is defined as the ratio between the received
power (in a matched load) and the radiated power. The coupling between two
parallel doublet antennas, spaced d apart is (— 18.46 — 20 log10(d/X))dB in the far
field. The general formula for the coupling between two antennas then becomes
(— 22 — 20 Iog10(<2/A) + Gx + G2)dB where Gx and G2 are the gains, referred to
isotropic, of the two antennas in the direction of one to the other. This must also
VHF and UHF antennas 853
take into account any skew or displacement between them. This formula relates to
free-space coupling and will be modified by the presence of reflecting surfaces pro-
viding transmission paths other than the direct one. A striking example of this is
given by Klimkiewicz50 for collinear antennas on a common metal mast. He shows
that the coupling between the antennas does not increase uniformly with distance,
but oscillates because each antenna illuminates the metal mast. As it is common
practice to mount many antennas for mobile services on a common mast it is
important to minimise their coupling in the manner indicated in this Chapter.
The problem of coupling between adjacent antennas reaches its most severe pro-
portions on aerospace vehicles, and considerable work on coupling analysis and
prediction has been reported in the IEEE EMC Symposium Records. Of the many
papers published on this subject, those of Siegel,81'82 Palas69 and Siakiewicz and
Adams79 are worthy of study.

± resonant at
T

Fig. 17.139 Dual-band monopole with single output

The formula quoted was for the far field for antennas intended for the same
frequency. The extent to which it can be used at close proximity depends on the
type of antennas and their arrangement, but it should give a reasonable approxi-
mation between parallel dipoles as close as one wavelength. When the antennas
are intended for different frequencies the effect of mismatch must be included;
it should be noted here that broad-band antennas may possess significant pass bands
at higher frequencies. It should also be noted that, although impedance characteris-
tics of antennas of the same type tend to be very similar within their design fre-
quency bands, this is not true outside these bands and it is unwise to rely on
measurements on single examples.
854 VHF and UHF antennas

17.8.2 Multiband antennas


A multiband antenna is one possessing acceptable electrical characteristics in two or
more separate frequency bands. The emphasis is on separated bands; whilst the log
periodic antenna may be considered as operating in a number of distinct bands
these overlap and this type of antenna would not be considered as multiband.

element C

insulator

metal sleeve A

Fig. 17.140 Two -frequency s/ee ve aerial

In general these antennas can be divided into two classes:


(a) Having a common physical structure and a single RF connector
(b) Sharing a common envelope but not necessarily a common structure and having
a separate RF connector for each frequency band.
The prototype of class (a) is shown in Fig. 17.139. The lower element (1)
resonates in the upper of two selected frequency bands, the resonant LC circuit
providing a very high impedance and effectively decoupling element (2) in this
frequency band. At the lower-frequency band the tuned circuit presents a series
inductance so that both elements are effective.
VHF and UHF antennas 855

A practical example of this type is shown in Fig. 17.140 with the dielectric sup-
port structure removed for clarity. It was designed by the author for use on high-
performance aircraft and may be mounted on the fuselage or inside a fin cap. It is
formally described as a top-loaded dual band sleeve antenna. Sleeves A and B
constitute a conventional broad-band sleeve monopole operating over the UHF
band 225-400MHz. The inner conductor of the coaxial feeder is connected to the
closed base of sleeve B, and the dimensions of the two sleeves are chosen to
optimise performance in this band, the total height being approximately 30 cm.

printed -circuit
board - ^ - _ _ _ _ / / // glass fibre
~ cover

glidepath
parasitic
notches

glidepath notch
* localiser
notch
NX \\

coaxial feed

Fig. 17.141 Two antennas in common surface

The rod within sleeve B forms with the inside of this sleeve a short-circuited
coaxial line having a high impedance at its open upper end over the UHF band
so that the horizontal top rod can be removed without significant effect. In the
VHF band 118-136 MHz this shorted line provides an inductance in series with the
top loading element whose dimensions are optimised for this band. The whole
antenna thus operates as a top-loaded sleeve for the VHF band. An important
point in the design of this class of antenna is that the upper band can be optimised
first and the components for the lower band added without significant disturbance
of the upper band performance.
As an example of the second class, Fig. 17.141 may be considered. This antenna,
for glidepath and localiser reception on aircraft, consists of two separate notch
856 VHF and UHF antennas

antennas in a common conducting surface. Because of the disparity in frequency


and the single pass-band characteristic of these notch antennas there is no
significant effect of one antenna on the other. This and the previous antenna are
described in more detail by Burberry.14

17.9 References

1 ADAMS, A. T., GREENOUGH, R. K., WALLENBERG, R. F., MENDELOVICZ, A., and


LUMJIAK, C: 'The quadrifilar helix antenna', IEEE Trans., AP-22,1974, pp. 173-178
2 ALBERTSEN, N. C , HANSEN, J. E., and JENSEN, N. E.: 'Numerical prediction of radi-
ation patterns for antennas mounted on spacecrafts'. IEE Conference Pub. 77, 1971,
pp.219-228
3 AWADALLA, K. H., and MACLEAN, T. S. M.: 'Input impedance of a monopole antenna
at the centre of a finite ground plane', IEEE Trans., AP-26,1978, pp. 244-248
4 BACH ANDERSON, J., and BALLING, P.: 'Admittance and radiation efficiency of the
human body in the resonance region', Proc. IEEE July 1972, pp. 900-901
5 BACH ANDERSON, J., and HANSEN, F.: 'Antennas for VHF/UHF personal radio; a
theoretical and experimental study of characteristics and performance', IEEE Trans.,
VT26, 1977, pp. 349-358
6 BARBANO, N.: 'The Aerodiscone antenna',Microwave /., Nov. 1966, pp. 57-62
7 BOSSE, H.: 'Rohrschlitzantennen mit horizontaler Richtwirkung', FTZ, (3), 1953, pp.
123-127
8 BOTHE, H.: 'In-flight measurement of aircraft antenna radiation patterns', AGARD,
1973, CP139, paper 35
9 BROWN, A. H., and STANIER, H. M.: 'Recent developments on VHF Ground-
Communication aerials for short distances', /. IEE, 94, Pt.IIIA, 1947, pp. 637-643
10 BROWN, G. H., and WOODWARD, O. M.: 'Experimentally-determined impedance charac-
teristics of cylindrical antennas', Proc. IRE, Apr. 1945, pp. 257-262
11 BRUECKMANN, H.: 'Improved wide-band VHF whip antenna', IEEE Trans., VC15,
1966,pp.25-29
12 BURBERRY, R. A.: 'Progress in aircraft aerials', Proc, IEE, 109, Pt. B, 1962, pp. 431-444
13 BURBERRY, R. A., and KELLY, W. A.: 'A review of helicopter aerial problems', IEE
Conf. Pub. 77,1971, pp. 77-82
14 BURBERRY, R. A.: 'The rationalisation of aircraft antennas', IEE Conf. Pub. 128,1975,
pp.204-209
15 CARREL, R. L.: 'The design of log-periodic dipole antennas', IRE Internat. Conv. Rec.
Vol.9,Pt. 1, 1961, pp. 61-75
16 CARTER, P. S.: 'Antenna arrays around cylinders', Proc. IRE, Dec. 1943, pp. 671-692
17 CARY, R. H. J.: 'The slot aerial and its application to aircraft', /. IEE., 99, Pt. Ill, 1952,
pp.197-210
18 CLAVIN, A., HUEBNER, D. A., and KILBURG, F. J., 'An improved element for use in
array antennas', IEEE Trans., AP-22,1974, pp. 521-526
19 COOPER, C. E.: 'An airborne blade antenna for 26-100 MHz', IEE Conf. Pub. 77,1971,
pp.54-59
20 COX, R. M., and RUPP, W. E.: 'Circularly polarised phased array antenna element', IEEE
Trans., AP-18, 1970, pp. 804-807
21 CRONE, W. J., and POPP, H.: 'Doppler VOR ground equipment', Electrical Commun. 43,
1968,pp. 143-149
22 CUMMING, W. A., and CORMIER, M.: 'Design data for small annular slot antennas',IEEE
Trans., AP-6, Apr. 1958, pp. 210-211
VHF and UHF antennas 857
23 DAVIDSON, A. L., and TURNEY, W. J.: 'Mobile antenna gain in the multipath environ-
ment at 900 MHz1, IEEE Trans., VT26,1977, pp. 345-348
24 DIFONZO, D. F.: 'Reduced size log-periodic antennas*, Microwave /., Dec. 1964, pp. 37-
42
25 EARP, C. W., and COOPER-JONES, D. L.: 'The practical evolution of the commutated
aerial direction finding system', Proc. IEE 105 Pt.B Suppl. 9, 1958, pp. 317-325
26 EHRENSPECK, H. W.: The backfire antenna, a new type of directional line source', Proc.
IRE, 48, 1960, pp. 109-110
27 GARFIELD, W. L., and NORBURY, F. T.: 'Instrument low-approach system and radio
altimeter for all-weather landings', Electrical Commun., 41, 1966, pp. 196-214
28 GOUGH, M. W.: 'Assessing reliable ranges in the VHF aeronautical band'. Symposium on
extended range VHF air/ground communications. International Aeradio Ltd., 1963,
pp.115-127
29 GRIFFITHS, R. M., and ROSINSKI, W.: 'The extension of wireless direction finding tech-
niques to very high frequencies for naval use', /. IEE, 94, Pt. IIIA, 1944, pp. 727-740
30 GUERTLER, R.: 'Impedance transformation in folded dipoles', /. Brit. IRE, 9, 1949, p.
344
31 HARRINGTON, R. F.: 'Matrix methods for field problems', Proc. IEEE, 55, 1967, pp.
136-149
32 HARRINGTON, R. F., and MAUTZ, J. R.: 'Straight wires with arbitrary excitation and
loading',IEEE Trans., AP-15, 1967, pp. 502-515
33 HARRINGTON, R. F.: 'Field computation by moment methods'. MacMillan, New York,
1968
34 IMBRIALE, W. A.: 'Application of the method of moments to thin wire elements and
arrays' in MITTRA, R. (Ed.): 'Numerical and asymptotic methods in electromagnetics'.
Chapt. 2, Springer Verlag, Berlin, 1975
35 JAMES, J. R., and BURROWS, R. M.: 'Resonance properties of dielectric-loaded short
unipoles', Electron. Lett., 9, 12 July 1973, pp. 300-302
36 JAMES, J. R., and HENDERSON, A.: 'Electrically short monopole antennas with dielec-
tric or ferrite coatings', Proc. IEE, 125, 1978, pp. 793-803
37 JAMES, J. R., and HENDERSON, A.: 'Investigation of electrically small VHF and UHF
cavity-type antennas', IEE Conf. Pub. 169, 1978, pp. 322-326
38 JASIK, H. (Ed.): 'Antenna engineering handbook', McGraw-Hill Book Co., 1961
39 JOHNSON, W. A., and ROTHE, P.: 'A wide-band circular slot radiator'. RAE Tech. Note
RAD453,1949
40 JOHNSON, W. A.: 'The notch aerial and some applications to aircraft radio installations',
Proc. IEE, 102, Pt.B, 1955, pp. 211-218
41 JONES, I. L.: 'Movements of the phase centre of ILS airborne localiser aerials on a Varsity
aircraft', RAE Technical Report 65241
42 JOSEPHSON, B.: 'The quarter-wave dipole', IRE Wescon. Conv. Rec, 1957, pp. 77-90
43 KANDOIAN, A. G.: 'Three new antenna types and their applications', Proc. IRE, Feb.
1946, pp. 70W-75W
44 KELLER, J. B.: 'Geometrical theory of diffraction', /. Optic. Soc. Am., 52, 1962, pp.
116-130
45 KILGUS, C. G.: 'Resonant quadrifilar helix design', Microwave J., Dec. 1970, pp. 49-55
46 KING, R.: 'Asymmetrically driven antennas and the sleeve dipole', Proc. IRE, 38, 1950,
pp.1154-1164
47 KING, H. E., and WONG, J. L.: 'A shallow ridged-cavity crossed slot antenna for the 240
to 400 MHz frequency range', IEEE Trans., AP-23, 1975, pp. 687-689
48 KING, H. E., and WONG, J. L.: 'Effects of a human body on a dipole antenna at 450 and
900 MHz\ IEEE Trans., AV-25, 1977, pp. 376-379
49 KITCHEN, F. A.: 'The design of an omnidirectional aerial system for the frequency range
225-400Mc/s,Proc. IEE, 98, Pt. Ill, 1951, pp. 409-415
858 VHF and UHF antennas

50 KLIMKIEWICZ: 'Compatibility of transmitting and receiving antennas on a common


mast', 1st EMC Symposium, Montreux, 20-22 May 1975, Paper 71
51 KNIGHT, P.: 'Methods of calculating the horizontal radiation patterns of dipole arrays
around a support mast', Proc. IEE, 105, Pt. B, 1958, pp. 548-554
52 KOUYOUMJIAN, R. G.: 'The geometrical theory of diffraction and its applications' in
MITTRA, R. (Ed.): 'Numerical and asymptotic methods in electromagnetics', Chap. 6,
Springer Verlag, Berlin, 1975
53 KRAUS, J. D.: 'Antennas', McGraw-Hill, New York, 1950
54 KRAUS, J. D.: 'Radio astronomy'. Chap. 6, McGraw-Hill, New York, 1967
55 KRUPKA, Z.: 'The effect of the human body on radiation properties of small-sized com-
munication systems', IEEE Trans., AP-16,1968, pp. 154-163
56 KUBINA, S. J., and PAVLASEK, T. J. F.: 'Numerical techniques for vehicle antennas'.
IEEE Conf. Paper 73254, Toronto, Ont.
57 KUECKEN, J. A.: 'Packset radio antenna measurements', IEEE International Conv. Rec.
Pt. 5,1966,pp. 261-270
58 LAMBERTS, K., and PUNGS, L.: 'Experimentelle Untersuchungen an Meter - und
Dezimeterwellen — Antennen fur breite Frequenzbander', FTZ, Issue 5, 1950, pp.
165-173
59 LEE, W. C. -Y.: 'Preliminary investigation of mobile radio signal fading using directional
antennas on the mobile unit', IEEE Trans., VC 15, 1966, pp. 8-15
60 LYON, J. A. M., and HIATT, R. E.: 'Electrically small antennas with loading materials and
with active elements'. Proc. E COM-ARO Workshop on Electrically Small Antennas,
US Army Electronics Command. Fort Monmouth NJ, Oct. 1976, pp. 205-210
61 MAHONEY, J.: 'Fincap communication aerials', IEE Conf. Pub. 77,1971, pp. 71-76
62 MAHONEY, J.: 'Upper L-Band telemetry aerials for rockets and missiles', AGARD CP
139,Paper 23,1973
63 MAYES, P. E., and CARREL, R. L.: 'Logarithmically periodic resonant-V arrays', IRE
Wescon Record, 1961
64 MEIER, A. S., and SUMMERS, W. P.: 'Measured impedance of vertical antennas over
finite ground planes', Proc, IRE, June 1949, pp. 609-616
65 MILLARD, G. H.: 'The evolution of antenna designs for u.h.f. relay stations', IEE Conf.
Pub. 169,1978, pp, 147-8
66 MILLER, E. K., et al:. 'Numerical analysis of aircraft antennas' in WAIT, J. R. (Ed.):
'Conf. on Environmental Effects on Antenna performance', Oct. 1969, pp. 55-58
67 MILLS, B. Y., and LITTLE, A. G.: 'A high resolution aerial system of a new type',
Australian J. Phys., 6, 1953, pp. 272-278
68 NAIL, J. J.: 'Designing discone antennas', Electronics, August 1953, pp. 167-169
69 PALAS, R. J.: 'Near field antenna-to-antenna coupling calculation procedure'. IEEE
13th EMC Symposium Rec, 1971, pp. 132-136
70 PAPAS, C. H., and KING, R.: 'Input impedance of wide-angle conical antennas fed by a
coaxial line', Proc. IRE, Nov. 1949, pp. 1269-1271
71 PLATTS, G. C : 'Provision of transmitting aerials at the main u.h.f. television stations in
the UK', Electron. & Power, June 1978, pp. 422-428
72 POPP, H.: 'Solid-state VOR: A new generation of omnirange navigation aids', Elect.
Commun., 44, 1969, pp. 322-329
73 PUGH, B., BAKER, D. G., and THOMAS, D. C: 'A sounding rocket omni-directional
antenna', IEE Conf. Pub. 77, 1971, pp. 1-6
74 RASHID, A.: 'A representation of cylindrical antennas for manpack installation', IEEE
Trans., AP-15, Sept. 1967, pp. 699-700
75 RICHMOND, J. H.: 'A wire grid model for scattering by conducting bodies',IEEE Trans.,
AP-14, Sept. 1966, pp. 782-786
76 RYLE, M.: 'A new radio interferometer and its application to the observation of weak
radio stars', Proc, R. Soc. 211 A, 1952, pp. 351-375
VHF and UHF antennas 859

77 SCHELKUNOFF, S. A., and FRIIS, H. T.: 'Antennas, theory and practice', John Wiley,
New York, 1952
78 SHNITKIN, H., and LEVY, S.: 'Getting maximum bandwidth with dipole antennas',
Electronics, 31 Aug. 1962, pp. 40-42
79 SIAKIEWICZ, K. R., and ADAMS, A. T.: 'Analysis and prediction of coupling between
collocated antennas', IEEE 14th EMC Symposium Rec, 1972, pp. 315-320
80 SIDFORD, M. J.: 'A radiating element giving circularly-polarised radiation over a large
solid angle', IEE Conf. Pub. 95, 1973, pp. 18-25
81 SIEGEL, M. D.: 'Aircraft antenna-coupled inteference analysis',. IEEE 11th EMC
Symposium Rec, 1969, pp. 85-90
82 SIEGEL, M. D.: 'Near-field antenna coupling on aerospace vehicles', IEEE 12th EMC
Symposium Rec, 1970, pp. 211-216
83 SINCLAIR, G.: 'The patterns of antennas located near cylinders of elliptical cross-section',
Proc. IRE, June 1951, pp. 660-668
84 SMITH, C. E.: 'Log-periodic antenna design handbook'. Smith Electronics Inc., Cleveland,
Ohio, 1966
85 SMITH, M. S.: 'Properties of dielectrically loaded antennas', Proc. IEE, 124, 1977, pp.
837-839
86 STIDHAM, J. R.: 'Experimental study of UHF mobile radio transmission using a directive
antenna', IEEE Trans., VC-15, 1966, pp. 16-24
87 THOMAS, R. L., and JOHNSON, R. B.: 'A flush cavity VOR/Localiser antenna', Micro-
wave J., Dec, 1966, pp. 66-70
88 WALKINSHAW, W.: 'Theoretical treatment of short Yagi aerials', /. IEE, 93, Pt. IIIA,
1946,pp.598-610
89 WANSELOW, R. D., and MILLIGAN, D. W.: 'A compact, low profile, transmission line
antenna - tunable over greater than octave bandwidth', IEEE Trans., AP-14, 1966, pp.
701-707
90 WEBSTER, R. E.: '20-70 Me Monopole antennas on ground-based vehicles', IRE Trans.,
AP-5,1957, pp. 363-368

General references on Aircraft Antennas


91 GRANGER, J. V. N., and BOLLJAHN, J. T.: 'Aircraft antennas', Proc. IRE, 43, 1955, pp.
533-550
92 'Radio antennas for aircraft and aerospace vehicles', BLACKBAND, W. T. (Ed.): Techni-
vision Services, Maidenhead, England, 1967
93 'Aerospace antennas', IEE Conf. Pub. 77, 1971
94 'Antennas for avionics', AGARD Conf. Pub. 139, 1973
95 'Antennas for aircraft and spacecraft', IEE Conf. Pub. 128, 1975
Chapter 18

Coaxial transmission lines


and components
W. T . Blackband

18.1 Introduction

In most radio systems the output of the transmitter and the input of the receiver
are at considerable distances from the terminals of the antennas with which they
work. Whenever this distance is an appreciable fraction of a wavelength, or greater,
then the satisfactory operation of the system demands that attention be paid to the
design of the connecting link.
In this Chapter an account will be given of the main types of transmission line
so that a sound choice can be made for any desired system. In many cases the sys-
tem will require the use of such devices as balancing units, power splitters, hybrids,
etc. These are grouped under the heading 'line components'. They are discussed in
this Chapter because many of them are built up from lengths of transmission line
and because without them neither transmission line nor antenna could function in
the desired manner.
The chief types of transmission line used by antenna engineers are the coaxial,
the open wire and the hollow waveguide. Waveguides have been discussed fully in
many specialised volumes and are not treated again here. Because of their extensive
use and wide range of application, coaxial lines are considered first and in some
detail; the open wire lines are then discussed briefly. The Section concludes with
data on other forms of transmission line which may be useful to the antenna engin-
eer. For convenience screened twin lines have been included at the end of the
section dealing with coaxial lines.

18.2 Coaxial transmission lines

The typical coaxial line is that shown in Fig. 18.1 where the conductors are a pair
of coaxial circular cylinders, and this type will be discussed first. For a circuit at
power frequencies where the current density could be taken as evenly distributed
across the conductor cross-section with 'go' and 'return' paths it is simple and con-
venient to look upon the line as two current-carrying conductors. It is clear that as
Coaxial Transmission Lines and Components 861
these currents are equal and opposite in magnitude and coaxial in path, their net
magnetic field outside the line is zero. It may be noted en passent that, if the con-
ductors are not coaxial through some defect in manufacture, there will be a result-
ant magnetic field and this constitutes one of the three mechanisms of unwanted
coupling to and from a coaxial line.

Fig. 18.1 Coaxial line


The broken circle in Fig. 18.1a shows the mean conductor diameter, j{d + D)

At higher frequencies it becomes more convenient to concentrate upon the field


rather than upon the conductors — that is one looks upon the transmission
phenomenon as the propagation of an electromagnetic field guided by the two con-
ductors. The currents then are necessary to satisfy the boundary conditions where
the fields end on the conductors. The field set up in a coaxial line bounded by per-
fectly conducting walls at low frequencies is transverse electromagnetic (TEM); that
is one in which both the electric and magnetic fields have components only in the
transverse plane. Such a field is shown in Fig. 18.Ib. If the walls are not perfectly
conducting the wavefront that is basically plane is modified near the conductors
where it bends back, as shown in Fig. 18.1c. This bending back is necessary so that
the Poynting vector can have a component into the conductors in order to provide
the energy flow from the field which feeds the resistive heating of the conductors.
This energy flow from the field must be considered as an EM wave propagating
radially into the conductors. Its attenuation at frequencies above 1 MHz is very
rapid; for instance with a copper sheet at 1 MHz the wave falls to 1/eth of its ampli-
tude in passing through a distance of 0.066 mm, or to 1.6 x 10~7 in traversing
1 mm. Consider the wave travelling outward through the outer conductor, it will be
much attenuated by the time it reaches the outside; what remains will be almost
completely reflected back into the conductor at the marked mismatch at the
copper-air interface. This results in a voltage gradient along the outside of the con-
ductor of F T /unit length. It is usual to assess the leakage of energy through the
outer conductor in terms of the 'transfer impedance' ZT which is defined as:
ZT = VT/l ohm/unit length (18.1)
862 Coaxial Transmission Lines and Components

The penetration of the wave into the conductor was discussed by Howe1 using this
concept of a rapidly attenuated wave flowing through a transmission line in a
copper medium. A derivation of ZT in terms of the constants of the conductor has
been given in detail by Schelkunoff.2
The two great advantages of the concept of transfer impedance are that it is not
difficult to measure for cables under laboratory test or as installed, and also it leads
in many cases to a simple calculation of the coupling between pairs of circuits.3 In
very many cases this coupling through ZT is the determining mechanism of leakage
out of (and into) coaxial lines, and for this reason it has been taken by the IEC as
the specified parameter in the standardisation of the efficacy of screening of flex-
ible coaxial cables.4
Because the screening of cables is of such importance in the complex and tightly
packed radio systems of the present day, the transfer impedance will be discussed in
more detail with particular attention to its variation with frequency. Referring to
Fig. 18.2, curve (a) shows the variation of ZT for a tubular conductor. It will be
seen that at low frequencies it approximates to the DC resistance per unit length,
which is shown by the broken line. This is clearly what one would expect because in
the case of a direct current the voltage drop along both the inside and outside of
the outer conductor would be the same and this would be related to the current
carried by the resistance per unit length.

DC resistance
/unit length

1.0 10 100
frequency MHz

Fig. 18.2 Variation of transfer impedance with frequency


a For a tubular conductor
b For a normal braid
c For an 'optimised' braid
PC resistance per unit length

As the frequency increases ZT decreases, slowly at first and then very rapidly so
that the leakage becomes quite negligible at the upper radio frequencies (the upper
part of the HF band, the VHF band and higher bands).
Coaxial Transmission Lines and Components 863
Because cables with tubular conductors are mechanically rigid, they are of
limited usefulness and in most installations the cables used have conductors formed
by bands of copper wire. Curve (b) shows the variation of ZT with frequency for a
typical braided conductor of the traditional design. It will be noted that with rise in
frequency there is at first a slight fall in ZT followed by a rise which becomes more
rapid until it increases linearly with frequency. From this it is clear that at the
upper frequencies braided cables have markedly greater leakages than those with
tubular conductors.
Recent studies by Homann5 and Tyni6 have shown that by using braids with
rather larger gaps between the braiding wires than had previously been considered
good practice, the braid can be 'optimised' with the result that ZT follows a curve
such as curve (c) of Fig. 18.2. The word 'optimised' which is generally used in order
to describe such braids is not well chosen. The 'optimisation' procedure aims to
balance the couplings through the braid by the mechanisms which give rise to two
voltages in antiphase. A truly optimised braid would have these exactly balanced
and a Z T of zero. However, because some deviations from the ideal construction are
inevitable the truly optimised cable withZy = 0 cannot be manufactured, and what
are known as 'optimised' cables are at best still far from the theoretically possible
perfection. It may be noted that the substitution of a cable with an optimised braid
for one with a traditional well-filled braid may typically reduce ZT by a factor of 3.
If this is done in both the interfered and interfering systems, then an improvement
of the order of 20 dB can be expected. (This figure will of course depend upon the
system frequency and the size of cable used.) Because of this improvement, 'opti-
mised' cables made to conform to IEC requirements forZ T should be used in instal-
lations where leakage must be minimised.
It is possible to reduce the leakage from a cable with a braided conductor by the
application of a second braid. With careful design the presence of this second braid
can reduce ZT by a factor of 30 or more, giving a reduction of radiated power of
30 dB. If the two braids are spaced by an intersheath ZT can be reduced by a fur-
ther factor of 10, resulting in a power decrease of 20 dB. For installations where
particularly low leakage is important a series of 'superscreened cables' has been
developed.7'8 The relative performances of these types of cables are set out in Fig.
18.3.
In order to complete the picture it should be noted that with cables having a
braided outer conductor a third mechanism of coupling is possible. This is by direct
capacitance coupling between the inner conductors by lines of electric field which
slip through the interstices of the braid. This can usually be ignored in single braid
cables, and is never of any consequence when a double braid is present.
The two characteristics of a coaxial line which are usually of prime importance
are the characteristic impedance Zo and the attenuation per unit length a. The
values of Z o have been standardised by IEC as 50 and 75 12 with some supplemen-
tary higher values (e.g. 100 and 12512) for low-capacitance cables. The 7512
standard was chosen because, for an airspaced cable with both conductors of the
same resistivity, it gives the minimum attenuation per unit length within a given
864 Coaxial Transmission Lines and Components

diameter. The 50 £2 value represents a compromise between the impedances which


would give minimum attenuation per unit length, maximum power-handling
capacity and maximum voltage rating.9 The 50 £2 cables are used in the vast majority
of cases. In the design of an antenna it may, however, be necessary to build a short
section of line to meet special requirements; the optimum proportioning of the con-
ductors when other properties such as voltage rating are dominant has been dis-
cussed.9' 10

Type of cable Zo,£2 D0D, 5 mm Zf, mS2/m


Single braid of normal design
URM43 or RG/58/U 50 2.95 300 1
URM67 or RG/213/U 50 7.25 300

Single braid of 'optimised' design to• IEC standard


URM43or RG/58/U 50 2.95 1002
URM67 or RG/213/U 50 7.25 100
URM57 75 7.25 100

Double braid in contact to IEC standard


UR91 orRG/214/U 50 7.25 32
URM43or RG/55/U 50 2.95 6
Double braid with intersheath
UR9 50 7.25 0.33
URM115 50 2.95 0.6

Superscreened cables
MM11/50 50 2.95 0.1 4
MM 12/50 50 7.25 0.03
Notes
1. Average values, very much larger values can occur
2. These would be the acceptance maxima which would apply if these cables were ordered to
meet IEC screening requirements
3. Estimated by author
4. Maximum specified in TRG 71181 (UKAEA)
5. At frequencies below about 4 to 5 GHz there is very little transfer of current from one
braiding wire to another. As a result, equal currents flow in the braid wires in contact with
the dielectric core and in those of the outer layer. For this reason it is usual to take the
effective diameter over the dielectric (DOD) as equal to the core diameter plus that of one
braiding wire. This applies to calculations of characteristic impedance, attenuation, cut-off
frequency etc.

Fig. 18.3 Transfer impedances of typical cables

Expressions for Z o and attenuation per unit length are given in Fig. 18.4. TheZ 0
depends only upon the ratio of the conductor diameters and \JJife. It is rare for the
relative magnetic permeability /i to be other than unity, and the only variable of
Coaxial Transmission Lines and Components 865
practical importance is the dielectric constant e. The expression for OL is more com-
plex. To begin with there are two terms: the first representing the resistive loss and
the second the loss in the dielectric between the conductors.

Conductor resistivity: p,
Dielectric constant; e
Relative magnetic permeability; ju
Power factor; p = tan 5
Frequency: f

Zo = 138 /-loglo(D/a0ohm
V e

a = 2.16

where px and p2 refer to the inner and outer conductors.


These equations can be combined to give:

a = 1.57 10"7 /- — I / l o g 1 0 - + 9 . 0 9 10~8/\/e tan 5 dB/metre


O / d
F ig. 18.4 ZQ and attenuation /unit-length of a coaxial line
D is the effective diameter over dielectric (DOD). See note 5, Fig. 18.3.

The resistive loss is proportional t o / , being controlled by skin effect. The second
expression for a shows the full dependance of the resistive term upon d and D. It
can be shown that, for a given value of D, a is a minimum in the case where px — p2,
if D/d = 3.59. For an airspaced line this gives a value of Z o = 76.7 £1. The standard
of 75 12 is chosen for low-loss air-spaced line in order to accomodate the inevitable
solid dielectric spacers which produce a slight lowering of the characteristic imped-
ance.
There are other criteria by which the optimum design may be judged other than
the diameter of the outer conductor. The cases of minimum dielectric and mini-
mum conductor cross-sections for a given attenuation have been studied.9'10
It is to be noted that both terms of the second expression for a are dependent
upon \Je. This means that by reducing the solid dielectric between the conductors
both the resistive loss and the dielectric loss are reduced. In fact, in the majority of
cases the reduction of the resistive loss is the more important benefit resulting from
866 Coaxial Transmission Lines and Components

the reduction of the effective dielectric constant of the line, by the replacement of a
continuous dielectric by one of the airspaced or semi-airspaced types.
It will be seen that the first term in the expression for ot is proportional to \Jf,
while the second varies with/. For this reason, while the copper losses dominate the
attenuation at the lower frequencies, if the frequency is increased sufficiently the
dielectric loss will become the more important.

100i

10-

o
o

00

10-

01-
1-0 10 100 1000 10,000
MHz
Fig. 18.5 Variation of attenuation with frequency
(a) For small solid polythene cable RG-58B/U
(b) For medium solid polythene cable RG-213/U
(c) For semi-airspaced cable of same DOD as ib)

In Fig. 18.5 are shown plots of attenuation per unit length against frequency for
some typical cables. Curve (a) relates to a small braided cable (UR43, RG-58/U). It
will be seen that at the low frequencies the slope approximates to that of a square-
root law and that it becomes steeper as the frequency increases and the linearly de-
pendent dielectric loss becomes more significant. Curve (b) is for a larger braided
cable (UR67, RG-213/U); its attenuation is lower because of its greater conductor
area, and the steepness is greater because the reduction of the resistive loss makes
that due to the dielectric relatively more important. Curve (c) shows the variation
for a semi-airspaced line of the same diameter as the solid dielectric line of curve
(b). The attenuation is markedly lower because:
(i) The continuous tube outer conductor has a lower resistance than the braid of
the second cable (UR67, RG-213/U).
(ii) The change from solid polythene to air-spaced dielectric reduces the effective e,
reducing both resistive and dielectric losses.
Coaxial Transmission Lines and Components 867

(iii) The above change reduces the power factor, and consequently the dielectric
loss.
It will be noted that the dielectric-loss term is not dependent upon the conductor
sizes. For a given value of tan 5 it is the same for large cables as well as small ones;
in fact, it is also independent of the cross-sectional geometry and applies to any
TEM system.

18.2.1 Frequency limitations


The useful upper limit of working frequency for a coaxial line can arise in two main
ways: by the increase in internal resonance or the setting up of waveguide modes.
The internal resonances occur when there are irregularities in the characteristic
impedance along the length of the line. These set up reflected waves returned to the
input, and so disturb the input impedance. Because the return wave is the resultant
of many reflections arising from different parts of the line, its resultant amplitude is
very sensitive to small changes in frequency, and it is the frequency-sensitive mis-
match which limits the usefulness of a line at high frequencies. Unless special care is
taken in the manufacture of flexible polythene cables the irregularities are likely to
be troublesome at about 1 GHz. Special cables to close tolerances are made and
tested for use at 3 and 10 GHz. The second limitation arises because at sufficiently
high frequencies a septate waveguide mode can propagate. The lower cutoff fre-
quency for this mode is that at which the mean conductor diameter \{D 4- d)
shown by the broken circle in Fig. 18.1a becomes a half wavelength long for a wave
travelling in the dielectric of the line. The above statement is only approximate, but
as it tends to predict a slightly lower frequency limit than does an exact analysis,
there is a safety factor in its use. The cut-off frequencies fc for this septate mode
for various cables are shown in Fig. 18.6.

Cable DOD fc, GHz Xc, cm


mm
UR94 1.05 50 90 0.33
UR95 1.60 50 62 0.48
UR43 2.95 50 31.8 0.94
UR67 7.25 50 13.0 2.30
UR57 7.25 75 14.8 2.03
UR74 17.3 50 5.7 5.26

Fig. 18.6 Cut-off frequencies for septate modes in polythene cables

The cables listed in this Figure all have solid polythene cores; they range in size
from the smallest to largest. It is useful to compare UR 61 and UR 57. These are
both the same diameter over the dielectric but the second cable being of Z o = 75 £2,
has a smaller inner conductor than the 5012 one; this results in a higher cut-off fre-
quency. The values of fc quoted have been calculated allowing for the thickness of
868 Coaxial Transmission Lines and Components

the braid wires, and therefore are not quite inversely proportional to the diameter
over the dielectric.

18.2.2 Voltage and thermal ratings


The power-handling capacity of a coaxial line can be limited by voltage breakdown
or overheating. In general, the former occurs with pulse systems, while it is only
with continuous signal systems that thermal breakdown becomes a problem.
The voltage breakdown in an air-spaced line occurs at the surface of the inner
conductor where the voltage gradient is greater than that at the outer conductor.
The working voltage V is related to the voltage gradient E by

F = \Ed\oge- (18.2)

For air-spaced lines, and also for semi-air-spaced lines with bare conductors, the
maximum safe voltage gradient ZT, is usually taken as 10kV(peak)/cm.
For cables with solid dielectric the voltage limitation is usually set by the break-
down of the air gap between the solid dielectric and the inner conductor, although
in some cables of low Z o the breakdown between the dielectric and a wire braided
outer conductor may set the limit. In general, the safe limit for E defined by the
airgaps is very much less than the intrinsic dielectric strength of the solid dielec-
tric.11
The thermal breakdown can take place when the temperature of the inner con-
ductor exceeds the safe limit. For polythene this is usually taken as 85°C and for
PTFE (Teflon or Fluon) it is taken as 250°C. The higher working temperature of
PTFE permits the manufacture of cables of considerably higher power-handling
capacities than their polythene equivalents. This is evident from Fig. 18.7.
The temperature rise of the inner conductor at a given, power level depends upon
the thermal resistance of the layers of the cable external to it. This has been dis-
cussed by Mildner12 whose methods are generally accepted for the calculation of
power ratings. It should be noted that the power ratings quoted in specifications
and cablemakers' lists may be based upon different ambient temperatures, and this
must be allowed for when comparing power-rating data from various sources. It is
also important to allow for any difference between the operating ambient tempera-
ture in a system under design and that chosen in the specification or list. The table
in Fig. 18.8 shows the correction factors which convert the power ratings calculated
at 55°C, 40°C and 30°C to various other ambients for cables with polythene insula-
tion.

18.2.3 Screened twin lines


A section through a screened line is shown in Fig. 18.9. The two identical inner
conductors are symmetrically placed and parallel to the axis of the outer conduc-
tor. The expressions forZ 0 and a apply to operation in the balanced mode in which
the currents in the two inner conductors are equal and opposite, so that there is no
net longitudinal current in the screen. For a given set of conductors, that is with d
Coaxial Transmission Lines and Components 869
and D fixed, Z o has a broad maximum when the spacing/diameter ratio h is equal to
0.486; for this reason it is usual to design screened twin cables with h close to this
value because then Z o is insensitive to the inevitable variations in cross-section

3000-

1000-

300-

100-

100 200 300 600 1000 3000


MHz

Fig. 18.7 Comparison of the power ratings of cables having cores of the same size but differ-
ing in material
PTFE core
Polythene core
UR102 and UR67 have DOD = 7.25 mm
UR108 and UR43 have DOD = 2.95 mm

Rating based on Ambient operating temperature


ambient 0°C 20 30 40 55 65 75 85
55°C UK services 2.85 2.17 1.83 1.50 1.00 0.67 0.33 0.0
40°C US services 1.89 1.67 1.22 1.00 0.67 0.44 0.22 0.0
30°C commercial 1.55 1.36 1.00 0.82 0.55 0.36 0.18 0.0

Fig. 18.8 Correcting factors for power ratings allowing for change in ambient temperature
(polythene insulation)

which occur during manufacture. A convenient and simple construction, based


upon twisting two similar insulated cores, results in a value of h — 0.500, and this
does not differ from the optimum by a significant amount.
It should be noted that the presence of the outer screen affects the Z o of the line
because it restricts the volume through which the field can extend. Thus for the
870 Coaxial Transmission Lines and Components
cable Duradio no. 11 the characteristic impedance is 93 £2, but if the screen were
removed (as in Duradio no. 20) Z o would increase to 120 £1. Twin cables also have
higher attenuations and lower power ratings than coaxial cables of the same size. If
the braid wires of a screened twin cable become corroded so that they do not make
good contact, then the screen fails to perform its function and, in the instance of
Duradio no. 11, the cable takes up a value of Z o somewhere between 93 and 120£2,
the precise value varying as thermal movements and mechanical vibrations change
the contact resistance between the wire braids. These effects do not and cannot
occur in coaxial cables;13 for this reason, when a flexible balanced line must be
used, this difficulty can be avoided if it is made up of a pair of coaxial braided
cables run in parallel, the input being between their inner conductors, with both
screens bounded together and earthed.

Fig. 18.9 Characteristics of a screened twin line


ohm

a = 2.16 10-\/f + 9.09 10"%/?tan 6 dB/metre

where g = d/2s h = s/D


Pj = resistivity of inner conductors,
p = resistivity of screen

In some applications it may be convenient to operate a twin line in the unbalan-


ced mode with both its inner conductors operated as one against the outer conduc-
tor. In this case, with the symbols of Fig. 18.9, the characteristic impedance is:

— h4)) ohm (18.3)

This is markedly lower than the characteristic impedance in the balanced mode.
Thus for Duradio no. 68, while the balanced Z o is 100 £2, that in the unbalanced
mode is only 19.6. The use of the twin cable in the unbalanced mode can give a
convenient low Z o line for matching purposes; however, because the characteristic
impedance is very sensitive to manufacturing tolerances and because the accept-
ance tests on the cables do not include a measurement of the unbalanced Zo it is
Coaxial Transmission Lines and Components 871
wise where a long production run is planned to form the low Z o line by paralleling a
number of coaxial cables, connecting their inner conductors at each end. Fig. 18.10
shows the characteristic impedances which can be obtained by connecting two or
three cables of 50 or 75 O in parallel.

Paralleled cables, O 75 + 75 75 + 50 50 + 50 75 + 75 + 50 75 + 50 + 50 50 + 50 + 50
ResultantZ 0 ,n 37.5 30.0 25.0 21.4 18.8 16.7

Fig. 18.10 Resultant characteristic impedances of paralleled cables

It is not necessary to give here the details of the construction and performance
of individual cables because these are readily available. There is an international
specification for radio-frequency cables — IEC96 issued in several parts.4 There are
also a number of UK specifications which are aligned with the IEC specification,
but having additional cables of types for which international specifications have not
yet been agreed. Thus the UK cables of the Uniradio and Duradio range are covered
by BS2316 and the US cables of the RG series are defined in MIL-C-17-D. In
addition most cablemakers issue their own lists which usually relate their products
where possible to the relevant items in the IEC, British Standard or US MIL docu-
ments.

18.3 Open-wire transmission lines

The most familiar of the open-wire lines are those used for telephones, but they are
also used for the feeding of HF antennas. Other systems of similar open construc-
tion are used in many antennas which are designed for installation on metal structures.
Open-wire lines have the great advantage that they are cheap compared with co-
axial cables large enough to carry the same power, and also they are easy for
experienced linesmen to install. However, their balance can be spoilt by the
presence of nearby metalwork and when this happens, they are liable to radiate or
to pick up unwanted signals. For these reasons they are usually confined to trans-
mitting systems and not used to carry very small signals which need protection
from interference.
Open-wire lines are normally air-spaced and so fx/e= 1; this term is therefore
omitted from the formulae given in this Section. From the expression for the
characteristic impedance given in Fig. 18.11 and the curve for Z o plotted in Fig.
18.12, it is seen that, when the wires are well spaced, that is 2y/d is greater than
about 50, Z o varies only slowly with the spacing. For this reason a characteristic im-
pedance of 600 H is much used for transmitter feeders because at this spacing the
inevitable small movement of the wires in the wind will not have a. significant effect
upon Z o and upon the impedance matching of the system. The rapid fall in Z o with
decrease in spacing for close wires (e.g. for Z o = 200 £2) is evident from Fig. 18.12.
872 Coaxial Transmission Lines and Components

The exact expression for Z o given in Fig. 18.11 involves the function
cosh"1(2j/(i).- This can be conveniently evaluated using the identity cosh"1* =
log^Oc + \lx2 — 1). For large values of x the approximation cosh"1* = loge2x can
be used; if x > 3.7 the approximation is within 1% and within \% for x > 5. The
values of cosh"1 for values of x up to 5.0 are tabulated in Fig. 18.13.

Fig. 18.11 Field near an open wire line


Z n = 120 cosh"11

or 120 locu ^ + / -f -

)
or approximately 276 log 10 is accurate to 1 % if — > 3.7
which
Yd')' ' d
but the error increases rapidly as the spacing decreases further.
d — conductor diameter
2y — axial spacing;

The power-handling capacity of an open-wire line is generally set by the voltage


at which corona starts. The table gives the values of the corona inception voltage
reported by R. A. Smith.14 These figures were determined experimentally using
pairs of wires at 9 in spacing. They are tabulated below.

Voltage Type of wire* Imperial wire Diameter


kV (RMS) (lb/mile, copper) gauge (in)
28 100 14 0.080
32 200 11 0.116
38 400 8 0.160
>42 600 6 0.192
Wiremen describe hard-drawn copper wire in terms of its weight in pounds per mile
Coaxial Transmission Lines and Components 873

It was found that Peek's law applies and that the voltage gradient at the conduc-
tor for the inception of corona is greater for fine than for thick wires, being pro-
portional to [1 + (0.301/r)], where r is the wire radius in centimetres.
In many cases it is necessary to fit corona rings to the spreader insulators, but
this results in an increased capacitive loading which increases the electrical length of

•8 4 0 0 -

40 50 60
spacing/diameter ratio }2y/d

Fig. 18.12 Characteristic impedance of a two-wire open line


Zo = 120 cosh"1 [2y/d)
* 276 log10 (4y/d) if 2y/d > 3.7

the feeder. The same source14 discusses this and quotes measured values for insula-
tor capacitance.
It has been observed that some outbreaks of corona are intermittent while
others are continuous. This is because, if the onset of corona makes the impedance
mismatch at the transmitter worse, then less power is delivered and the corona is
extinguished; in other circumstances the mismatch may be reduced and the corona
become firmly established.
The lines of electric and magnetic field between two conductors of circular
cross-section are shown in Fig. 18.11. The importance of the lines of force drawn
874 Coaxial Transmission Lines and Components
there is that they can lead to further information about associated forms of trans-
mission line. For example the line AA', which is an axis of symmetry and an equi-
potential surface, could be occupied by a perfectly conducting sheet without
disturbing the fields. Because the fields either side of the sheet are mirror images
they represent the same voltage difference; hence the impedance between either
wire and the earth plane must be

Z o = 60 cosh"1 — (18.4)
d
This, then, is the characteristic impedance of a wire of diameter d parallel to a
plane conducting sheet at a distance y from the plane.

X cosh lx X cosh lx X cosh lx X coshlx


1.00 0.0000 2.00 1.3170 3.00 1.7627 4.00 2.0634
1.05 0.3149 2.05 1.3454 3.05 1.7803 4.05 2.0763
1.10 0.4436 2.10 1.3729 3.10 1.7975 4.10 2.0889
1.15 0.5411 2.15 1.3995 3.15 1.8143 4.15 2.1014
1.20 0.6224 2.20 1.4254 3.20 1.8309 4.20 2.1137
1.25 0.6931 2.25 1.4506 3.25 1.8472 4.25 2.1259
1.30 0.7564 2.30 1.4750 3.30 1.8633 4.30 2.1380
1.35 0.8140 2.35 1.4989 3.35 1.8790 4.35 2.1498
1.40 0.8670 2.40 1.5221 3.40 1.8946 4.40 2.1616
1.45 0.9163 2.45 1.5447 3.45 1.9098 4.45 2.1732
1.50 0.9624 2.50 1.5668 3.50 1.9248 4.50 2.1846
1.55 1.0059 2.55 1.5884 3.55 1.9396 4.55 2.1960
1.60 1.0470 2.60 1.6094 3.60 1.9542 4.60 2.2072
1.65 1.0860 2.65 1.6300 3.65 1.9686 4.65 2.2182
1.70 1.1232 2.70 1.6502 3.70 1.9827 4.70 2.2292
1.75 1.1588 2.75 1.6699 3.75 1.9966 4.75 2.2400
1.80 1.1929 2.80 1.6892 3.80 2.0104 4.80 2.2507
1.85 1.2257 2.85 1.7082 3.85 2.0239 4.85 2.2613
1.90 1.2572 2.90 1.7267 3.90 2.0373 4.90 2.2718
1.95 1.2876 2.95 1.7449 3.95 2.0504 4.95 2.2822
2.00 1.3170 3.00 1.7627 4.00 2.0634 5.00 2.2924

Fig. 18.13 Table ofcosh^x

We have thus derived an expression for another line from a knowledge of the
field pattern of a well known line. This process of introducing an imaginary per-
fectly conducting surface to follow the shape of an equipotential surface is of
general application. Thus a conducting cylindrical sheet could be placed over one of
the equipotential circles surrounding the upper conductor of Fig. 18.11 so as to
Coaxial Transmission Lines and Components 875

form a pair of lines of dissimilar diameters, as shown in Fig. 18.14. It is clear that
the Z o of this pair would be the sum of the characteristic impedances of each of
the lines formed by the conductors and an imaginary conducting plane at AA'. The
ZQ of each line is calculable using the above equation. This is simple if the position
of AA' is known; however, given only the diameters of the two conductors, D and
d, and their axial separation, 2y9 it is necessary to locate the position of the equi-
potential plane. In order to do this use is made of the properties of the orthogonal
circles which define the lines of electric field and equipotential surfaces.

Fig. 18.14 Characteristic impedance of a pair of wires of differing diameters

For the equipotential circles it is known that the product of the two intercepts
with the normal to A A' is equal to the square of the unit distance of the grid. As
this unit distance is the same for both conductors, referring to Fig. 18.14, one can
equate these products:
(yi+^OCVi-rO = (y2 + r2)(y2-r2) (18.5)
The other known dimension s is introduced through the equation
s = yx+y2 (18.6)
Solving for yx and y2

2s
(18.7)

2s
With the position of AA' known, the characteristic impedances Z0l and Z^ of
876 Coaxial Transmission Lines and Components
the cylinders with AA' can be calculated. Their sum Z 01 4- Z&, is the characteristic
impedance of the line formed by the two conductors:

+ = 601 cosh (18.8)


D
It is worth noting that their difference Z^ — ZOi is the characteristic impedance
of the eccentric line which would be formed by folding Fig. 18.14 about AA' so
that the smaller circle lay within the larger. Fig. 18.15 shows the resulting arrange-
ment. The relative position of the circles is defined by their axial spacing o. In order
to calculate Zo between the conductors it is necessary to locate AA'. This is done
by equating the products of the ordinates of the intersections of each circle and
OB; hence
and as o = O1—J2)
r\-
(18.9)
2o 2a

Fig. 18.15 Pair of eccentric circular conductors

The ZQ of the region between the circular conductors is equal to that between
the circle d and AA' less that between the circle D and AA'. Hence the Z o between
the circles is given by

(18.10)

The expression is exact even for the closest spacings of the circles; however, if
the eccentricity is small, the following expression quoted by Schelkunoff2 will be
sufficiently accurate for most purposes;

Q = ZQ11 — (18.11)
Coaxial Transmission Lines and Components 877
where Z o = characteristic impedance of the line if concentric, e = eccentricity
(interaxial-separation/radius-outer-conductor), and JC = D/d.
Schelkunoff also gives an expression for the attenuation of the eccentric line;
/ 2e2 e2x2 \
1 8 1 2
+ + ( >
where a = attenuation of the corresponding concentric line.
No indication is given of the limits of eccentricity for which the approximation
is reliable except that it is recommended for 'nearly coaxial' lines.
The 2-wire open lines have two sources of weakness: first because of their high
characteristic impedance they are more susceptable to stray couplings and also their
working voltages are higher than is the case for lines of lower impedance; secondly,
because the energy of the guided wave is spread at significant densities over a con-
siderable area about the line axis the chances of interference are relatively high.
Various forms of open-wire lines using three or more wires have been developed
in order to give a reduced Z o without the need for critical wire spacing and to give
some increase in power-handling capacity. Multiwire lines may be either sym-
metrical and therefore balanced, or they may be unbalanced, with one set of con-
ductors partially screening the other, thereby reducing interference by reducing the
field at points not in the immediate vicinity of the line.
Expressions for the Z o of various types of open-wire line are given in Figs. 18.16
and 18.17. These all assume that the wire spacings are much larger than the wire
diameters. The validity of these expressions at close spacings is discussed in a later
Section.
In Fig. 18.16a is shown a group of three wires with the two outer ones connec-
ted together. These two can be considered as a very degenerate form of coaxial
screen.
In Fig. 18.16Z? is shown the field around a set of four wires positioned at the
corners of a square and driven so that each of a diagonal pair are at the same volt-
age, the signal voltage being applied between the opposed pairs, it is therefore a
balanced line. It will be seen that there are two planes of symmetry AA' and BB'.
These are equipotential surfaces because they are normal to the lines of electric
field. Hence the system could be halved by a conducting plane put through AA'.
This is shown in Fig. 18.16c and it represents a balanced 2-wire line above a con-
ducting plane such as the earth. It is clear that Z o is halved.
In the same way the system of Fig. 18.16& could be quartered by conducting
planes lying in AA' and BB'. One of these quarters is shown in Fig. 18.16J; it is a
single conductor placed in a conducting right-angled corner. Because the four quar-
ters of Fig. 18.16Z? are connected in series-parallel, the Z o of each quarter is the
same as that of all four taken together; and for this reason the expression for Z o
given in Fig. 18.16<i is the same as that in Fig. 18.16&.
The formulae quoted for Z o in Fig. 18.16& and c are among those given by
McLean and Bolt,15 who also present data on other open-wire lines and the effects
of nearby conducting planes.
878 Coaxial Transmission Lines and Components

o oo
S -I

A--H H--A

T o o
h

2.

F ig. 18.16 Some open-wire fines


(a) 3-wire fine
All wires of diameter d. The voltage is applied between the inner and outer
wires
Z
o =
{b) 4-wire line
All wires of diameter d at corners of a rectangle of sides sx and s2. The voltage
is applied between diagonally opposed pairs

Z o = 138log K

\fSi = s2 =s Zo =

(c) Balanced two-wire line above earth


Both wires of diameter c/ spaced 5 from each other and h from earth.

Zo =
(d) Wire in corner
The wire of diameter d is spaced at distances sl/2 and s2l2 from the walls of a
conducting right-angled corner

Zo = 138log l 0
Coaxial Transmission Lines and Components 879

+ +

+ +
c
s ; 0
4- +
-0-

-fr:
4- +1

F ig. 18.17 Z o o f some open-wire lines


(a) 4-wire line
This differs from the 4-wire line of Fig. 18.15 in that adjacent wires are at the
same potential
125,

For a square, st = s2 and Z o = 138 logj


'\d~)
(b) 2 wires in parallel above ground

(c) 5-wire line


The four outer wires form a crude screen around the central wire
Zo - 173log l 0 M.07~
(d) 6-wire line
The four outer wires form a screen around the centre pair which are run in
parallel

= 276
880 Coaxial Transmission Lines and Components

The characteristic impedances of some other open-wire lines are given in Fig.
18.17. The solution of the case of two wires in parallel above ground shown in Fig.
18.17& is derived from Fig. 18.17a by placing a conducting plane in the equipoten-
tial surface midway between the pairs of wires. The Zo of each pair of paralleled
wires taken against the plane is half that of the system of Fig. 18.17a. The line with
two wires in parallel is not often used by design although it may arise inadvertently
when the feed to a balanced line is out of balance.
Figs. 18.17c and 18.17J show examples of unbalanced lines in which one set of
conductors is partially screened by the other; they may be regarded as rudimentary
coaxial transmission lines.

18.4 Miscellaneous parallel transmission lines

This Section considers various types of transmission line of uniform cross-section.


To simplify the discussion they are assumed to be air spaced; the term v£t/e~has
therefore been omitted from all formulae.
The coaxial line with confocal elliptical conductors shown in Fig. 18.18 has a
simple and exact expression for Z o

Zo = 60 cosh" (18.13)

Fig. 18.18 Cross section of confocal elliptical conductors

Shebes18has derived an approximate equation valid if the eccentricity is small

* - » » * •f
It will be noted that the first term is that of the corresponding line with circular
conductors and the second term represents the effect of distorting the line from a
circular cross-section.
Omar and Miller19 extend this analysis on a semi-empirical basis to line cross-
sections formed by rectangles whose major and minor axes define confocal ellipses.
This is extended to other rectangular cross-sections giving an accuracy in general of
better than 2%. A circular inner conductor within a square-section outer conductor
is shown in Fig. 18.19. The expression given by Schelkunoff20 for Z o is that for a
Coaxial Transmission Lines and Components 881
normal coaxial line with the addition of the factor 1.078. This factor in effect
replaces the square by a circle of diameter 1.078 times that of the inscribed circle.
The effect of the factor is to add 4.51 £1 to the value of Z o calculated for a circle of
diameter D. The reason why there is a limitation on the value of D/d is that, for a
very large inner conductor, the uniform distribution of charge on this conductor,
which would occur if the outer were circular, would be seriously distorted by the
proximity of the mid points of the sides of the square in which regions the charges
would concentrate. An expression for the Z o of a circle within a square applicable
to all diameters of inner conductor has been derived by Parzen.21

Fig. 18.19 Conductors in square and semi-circular tubes


(a) For D > d
Zo = 138iog l0
{b) For d < D,s

where, as in Fig. 18.9, h = s/D and g = d/2s

The second illustration in Fig. 18.19 shows a single conductor in a semicircular


tube. The expression for Z o is derived from Fig. 18.9 where it is seen that there is a
diametrical equipotential plane. If this is filled by a conductor, the system of Fig.
18.195 is formed in each half; hence the Z o of the whole line lies half in each semi-
circle, and the multiplying factor of 216 becomes 138.

18.4.1 Shapes of conductor cross-sections


Most of the expressions for Z o already quoted have related to transmission lines
having circular cylindrical conductors; sometimes other sections are used. Thus it is
quite usual to use stranded conductors in place of the more rigid solid wires, and
hollow braids instead of tubes. The effects of stranding have been studied by
Meyers22 whose data are the basis of Fig. 18.20. Values are quoted there for those
numbers of strands which are commonly used because they lie neatly within a
circle. As far as Z o is concerned, the important parameter is the number of strands
in the outer layer and this is tabulated in the second row. The effective diameter
De is the diameter of a single conductor giving the same Z o as that of the stranded
conductor of overall diameter D o , i.e.
De = kxDo
Because the effective diameter is less than the overall diameter, Z o will be greater
882 Coaxial Transmission Lines and Components

by AZQ compared with that of a line with a single wire inner, where

Z o = 1381og10*i (18.14)

Values of AZ 0 calculated for an air-filled system are given in the bottom row.

Total number of strands 3 7 12 19 27 37


Number in outer layer 3 6 9 12 15 18

*i 0.871 0.939 0.957 0.970 0.976 0.980


AZ0 8.30 3.76 2.48 1.83 1.47 1.22

Fig. 18.20 Effect of stranding upon Zo

Applying this to a balanced twin line it is seen that the use of 7-strand wire
would add 3.7612 for each conductor, i.e. 7.52 O for the whole line. It is clear that
this would also be the increase for the 4-wire lines of Figs. 18.16 and 18.17.
It is sometimes convenient to have the inner conductor of other than circular
cross-section, in which case it is necessary to determine the radius of the equivalent
circle. Three shapes and their equivalents are shown in Fig. 18.21.

T
a

h—b—H
, = 0.5a De= 1.18a 0 e —0.40 ( a + 6 )
Fig. 18.21 Diameters of circles equivalent to simple shapes

The values quoted for the equivalent circles are derived on the assumption that
the spacing of the conductors is large compared with the dimensions a and b. In
such circumstances the current distribution will not change as the conductor is rota-
ted about its axis and will be that for the conductor in free space.
The derivation of the equivalent circle for the thin ribbon or strip is published
by Anderson23 as a demonstration of the conformal method of computing charac-
teristic impedance.
It will be seen that the equivalent diameter for the field outside the square
quoted above, 1.18c, differs from the value when the field is inside the square,
1.078, shown in Fig. 18.19a. The value for the angle is a good approximation when
a and b are nearly equal but if b became very small, the equivalent diameter would
increase and approach the value of 0.5Ckz applicable to the ribbon. These consider-
ations do at least define the limits between which the equivalent diameters for
intermediate bja ratios must lie.
Coaxial Transmission Lines and Components 883
18.4.2 Conductors in enclosures
In some cases the field around a conductor will be limited in extent by parallel con-
ducting planes. The simple case of a conductor parallel to a plane is considered in
the discussion of Fig. 18.11 and those of series and paralleled conductors in Figs.
18.16c and lS.llb. A conductor in a square is shown in Fig. 18.19a and one in an
angle in Fig. \S.\6d.
Other arrangements are shown together with expressions for their characteristic
impedances in Fig. 18.22. The expression for the conductor sandwiched between
two parallel sheets is shown in Fig. 18.22a. The expression for Z o is quoted by
Jasik.16 It will be noted that the second term within the squared bracket which cor-
rects the first term (which by itself is often quoted as an approximate formula) is
surprisingly small, being no more than a correction of 0.58% for d/D = \ .
Fig. 18.22& shows an arrangement of balanced twin conductors between two
planes, in which Z o is twice that for the previous arrangement. The reason for this
can be seen from Fig. 18.22Z?, where the broken line is an equipotential surface. If a
conducting sheet were placed in this position the field pattern would not be
disturbed but the transmission line would behave as two lines of the type shown in
Fig. 18.22a, and would have twice the impedance of the line of Fig. 18.22a.
The unbalanced line of Fig. 18.22c and the balanced line of Fig. \%22d are like-
wise linked in Z o . The dotted line of Fig. \%22d marks an equipotential plane mid-
way between the conductors. If it were occupied by a conducting sheet, one would
have a line of the type of Fig. 18.22c on each side of A A'. Hence the Z o of Fig.
18.22^ must be twice that of Fig. 18.22c.
The expression for Z o given in Fig. 18.22c is applicable if the diameter of the
central wire is small. A detailed analysis by Chisholm24 has led to expressions which
are accurate for large-diameter wires and which can allow for displacements of the
wire from the central line. Corresponding comments apply to the expression given
in Fig. l&.22d which is derived from the preceding case.

18.4.3 Lines with strip or ribbon conductors


These lines are based upon conductors for which the thickness is small compared
with their other dimensions. The term 'strip' is used here because it is well estab-
lished, although 'ribbon' is more appropriate because it emphasises the uniformity
of width which is a characteristic of these conductors.
The general problems of calculating the Z o of strip conductors has been dis-
cussed by Magus and Oberhettinger.25 Such strips may lie in a plane or follow a sur-
face which is curved laterally, the typical one considered is of a strip with a sur-
rounding outer conductor.
The method of computation is to use a conformal transformation and Green's
function. In the general case it is necessary to use elliptic integrals in order to evalu-
ate Z o , but in certain special cases approximate formulae can be derived. The data
in Fig. 18.23 are based upon Table 1 of the Magus paper.25 Approximate formulae
for Z o are given for the cases where d/b < 1 or d/b > 1; for intermediate values of
d/b the above authors have published curves which assist in the evaluation of the
884 Coaxial Transmission Lines and Components
relevant expressions. It must be remembered that all the formulae are derived on
the assumption that the strip is of negligible thickness; that is to say, all currents
flow on the broad surfaces and none on the edges. In many cases this neglect of the
edge currents is justified.

2D D -

1
2h
I
I
I
-4-
Fig. 18.22 Characteristic impedances of some wires within enclosures
4
F "I

4D
0-0367 (tf/P)4 1
-
-
97fi
276 J
(c)Z0 = ^tanh^ cT < D, ft
ITU
ITU U
U
for an expression accurate for large values of d see Chisholm24
= 2276log
{d)Z00 = 7 6 l o g 1 10 0^§tanh^
§tanh^ d<Dfh
for an expression accurate forr large values double the expression given by
Chisholm24 for the case of (c) above

Considering Fig. 18.23; case a, when d/b < 1, can be regarded as a slot in an infi-
nite sheet; this is sometimes used as a 'slot line' feeder, but more frequently as the
basis of resonant slot antennas. Fig. 18.23& shows the two extreme cases of d/b < 1
and d/bP- 1. The intermediate cases have been considered by Morton.26 In Fig.
18.24 are shown three curves relating Zo with d/b. The left-hand one applies when
Coaxial Transmission Lines and Components 885

Line geometry Approximate formulae forZ 0


d/b<\ d/b> 1

257
276 log 10 4(2 +dlb)
log lo 4(1+26/c0

d/b<\ d/b>\
d
L 377 d/b 276log10 4 -
\ b

157T 2 (Bd
138logl0|--
0.693 + ^

d/b- 1
64.3 (Sd
138logl0 - ~
log io \n b

d/b> 1
64.3
138 Iog10( 2.16-
3.06 \ b

\-b/d
-d—H

d/b> 1
64.3
138logl0 2~
b
\-b/d

_ b— d/b>\
d
/2 d
4d
188- 138logl0
_ I b

Fig. 18.23 Z o of some strip conductor fines


886 Coaxial Transmission Lines and Components
d/b is very small. The other solid curve represents the equation
Z o = 276log1Q4d/b (18.15)
which applies when d/b is large. The dotted curve departs upwards from the right-
hand curve and shows values of Zo based on Morton's data. It is clear that the
broken curve must be tangential to the left-hand curve at the origin, but Morton's
data do not extend to very small values of d/b. The difference between the broken
and solid curves can be looked upon as a correction factor which is plotted in Fig.
18.25. It should be noted that the ordinates of this correction curve on the left
where it shoots upward are not defined accurately by Morton's data.

400- Z =276 log 4d/b


0 lU

2 300-
<u
"8

200-

100-

3 4
the spacing to width ratio d/b

Fig. 18.24 Z o of parallel strip transmission line


Zo based on Morton's data

Returning to Fig. 18.23 it will be noted that, when d/b > 1 for cases (c) and (d),
Zo is the same; this illustrates the point made previously that, when a non-circular
conductor is well away from adjacent conductors, it can be rotated about its axis
without affecting Z o .
In case (e) it will be noticed that, for d/b > 1, the multiplying factor 2.16 is the
factor 1.078 given for a circular conductor in a square in Fig. 18.19a divided by
that relating the width of a strip to the equivalent diameter from Fig. 18.21.
Example (g) of Fig. 18.23 has been added to the six cases given by Magus and
Coaxial Transmission Lines and Components 887
25
Oberhettinger by imagining a plane conductor, to be placed between the strips of
case (b) in order to derive an expression for Z o for the unbalanced line formed by a
strip parallel to a large conducting plane.

u-

13-

E 1-2-

1-1-

1-0-
1 2 3 4 5 6
d/b

Fig. 18.25 Correction factor by which Z o = 276log104d/b is to be multiplied in order to give


Morton's exact value

18.5 Conical transmission lines

All the transmission lines considered so far have had a uniform cross-section.
Conical lines have a varying cross-section but maintain a constant characteristic
impedance and therefore waves can travel along them without impedance mis-
match. The lines are again considered to be air-spaced and the term y/jjje omitted
from all the formulae for characteristic impedance.
Two typical examples of conical lines are illustrated in Fig. 18.26; the left-hand
line is the conical counterpart of the coaxial line and the right-hand one is the
counterpart of the balanced twin open-wire line. It is evident that the wavefront in
these lines will not be plane but a section of a sphere, and, further, will expand
keeping its shape constant.
Lengths of coaxial conical line have been used for many years as a means of
joining one coaxial line to another of lesser diameter without the impedance
888 Coaxial Transmission Lines and Components

discontinuity which would accompany any sudden steps in the diameters of the
conductors. With the aid of stereographic projection it is possible to derive the
characteristic impedance of a given conical line from that of a corresponding uni-
form line and vice versa. This technique was described by my onetime colleague Dr
P. G. Rohde in an unpublished internal note; because it is not well known in the
antenna literature the use of stereographic projection in the theory of conical lines
is given here in some detail, together with examples of its use in the determination
of the characteristic impedance of related uniform lines.

9j angle
between axes

8 half angle
of cones

Fig. 18.26 Characteristic impedance of conical lines


tan-^
(a)Z0 = 138log l0

(b)Z0 = 120 cosh

or approximately = 276 log^

In Fig. 18.27 is shown a pair of cones with apices at 0, the centre of a sphere.
These form a conical transmission line whose intercepts on the sphere are A and B.
Consider a tangent plane EFGH touching the sphere at Q. Then the projection of A
and B from P, the other end of the diameter through Q, giving A' and B' is a stereo-
graphic projection.
We now introduce the theorem that if a pair of conductors is erected on A' and
B' normal to the tangent plane, then the characteristic impedance of this pair is the
same as that of a conical pair. This follows from the discussion of spherical waves
on cones by Schelkunoff.27
In order to illustrate the method Fig. 18.28 and 18.29 have been prepared to
show the projections used in order to derive the characteristic impedances of the
two conical lines shown in Fig. 18.26. Consider first Fig. 18.28, which shows a
coaxial conical line consisting of two conductors, the inner one of half angle dx and
the outer of half angle 62. The conducting parts are shown shaded. The circle is of
unit diameter. The circles are projected from P onto the tangent plane at Q. The
lines of projection are shown dotted. The circles on the sphere project to give two
Coaxial Transmission Lines and Components 889
circles on the plane both centered on Q, one of radius QA and the other of radius
QB. We consider a coaxial line erected upon the circles; its Z o will be equal to

1381ogJ^H (18.16)

Fig. 18.27 Projection of a conical line onto a tangent plane

In order to express this in terms of the angles 6t and 62, which define the conical
line, it should be noted that the angle subtended at P by any arc of the circle is
half that subtended at O. Consequently
QA = tan B2 /2 and QB = t a n ^ / 2
Hence
[tanflx/2
1381og10 (18.17)
{tm62/2
By similar reasoning it is seen that the twin line erected upon the circles AA' and
BB' of Fig. 18.29 has a characteristic impedance

Z o = 120 cosh"1 (18.18)

and because
y = tan 0i/2
and
d - tan — tan
890 Coaxial Transmission Lines and Components
the Z o of the twin conical line is

Z o = 120 cosh"1 (18.19)


tan ( — ) - tan —)

Fig. 18.28 Stereographic projection of coaxial conical transmission line


dx — half angle of outer cone
02 — half angle of inner cone

Fig. 18.29 Stereographic projection of twin conical transmission line


Sx = half angle between axes
0 2 = half angle of cones

Taking the usual simplification for 9X much greater than 62, the expression
becomes

Z o = 2761og10 (18.20)

It may be noted that the reason why the dotted cones of projection from P can
Coaxial Transmission Lines and Components 891
intersect both the sphere and the tangent plane in circles, is that the cone of projec-
tion is elliptical in cross-section and the circular cap and tangent plane both make
the same angle with the axis of the projection cone from P. It is an important
property of stereographic projection that any circle on the sphere is projected as a
circle on the tangent plane.
In Fig. 18.30 are shown two conical lines which are of special use in antenna
theory: the cone normal to a plane conductor which is introduced in the study of
the monopole excited against a plane, and the two identical cones with a common
axis (which mathematicians would describe as only one cone) on which dipole
theory is based.25
The expression for the Z o of the former is derived from the equation of Fig.
18.27# by putting dl = 7T/2; the corresponding expression for the two cones is
obtained by doubling this. It is to be noticed that the Z o of a cone inclined to a
conducting plane will be half of that shown in Fig. 18.27& for the twin conical line.

Fig. 18.30 Z o of two simple conical lines


(a) Cone normal to plane
Z o = 138 log 10 cot 0/2
(b) Two cones with common axis
Z o = 276 log lo cot0/2

If the cones of Fig. IS30b were rolled flat one would have the conical strip of
Fig. 18.31a. A half of this, driven against a conducting plane, is shown in elevation
in Fig. 18.3 \b. We now derive an expression for the Z o of the conical strip starting
with the arrangement shown in Fig. 18.31c. This shows the intersections of the con-
ducting surfaces of the half strip system with the sphere. The strip intersects the
sphere along an arc subtending 20 at the centre. This is projected from P as a
straight line BB' in the tangent plane. A narrow strip of uniform width is shown
erected upon BB'. The plane conductor intersects the sphere around the equator.
This cut projects as the circle AA' in the tangent plane. We have thus the arrange-
ment of a strip within a circular conductor as shown in Fig. 18.3 Id.
892 Coaxial Transmission Lines and Components

The ZQ of a strip within a circle has been previously noted (Fig. 18.24/) to be
equal to that of a circular conductor of diameter half the width of the strip within
the same circle, or
Id
Zo = 1381og10— (18.21)
b

A' d

Fig. 18.31 Conical strips and plane transmission line

provided d/b > 1. This is, referring to Fig. 18.


/2x AA'\
Zo = 1381og10 (18.22)
\ BB' /
BB' = 2 tan0/2 and AA' = 2.
(the sphere diameter is of unit length). Hence

Zo = 1381og10(2cot<9/2) (18.23)
For a conical strip on a plane, we double this in order to obtain
ZQ = 276 log 10(2 cot0/2) (18.24)
for the conical strip of Fig. 18.3 Iff.
Coaxial Transmission Lines and Components 893
Now the relationship between the width of the strip and the diameter of the
equivalent circular conductor will only apply if the width is small compared with
the diameter of the outer conductor; that is in the present case if 6 is small.
An expression for other values of 6 is derived in the following way; consider the
strip line of Fig. 18.32a having half angle 0, and the complementary one having half
angle f— 6 shown in Fig. IS32b. These are not only complementary geometrically;
they support complementary fields, and therefore from Babinet's principle their
characteristic impedances are related in air by the formula
ZoZo = T/2/4 where n = 1207T - 377.

Fig. 18.32 Two strip transmission fines forming complementary screens


If 6 < 45°, then the characteristic impedances are (fora) Z o = 276 iog lo 2cot0/2,
128.6
and (for/?) Z*Q =
• o.... . (2
( cot

Hence for 6 near to TT/2,


128.6
Z'o = r?2/4Z0 = - (18.25)

We thus have accurate expressions for 6 small and near to 7r/2; an exact value can
be obtained for the case 0 = n/4 because then the two lines are identical, and there-
fore
Zo = Zo = ??/2 = 188.5 O (18.26)
It will be found that at 6 = n/4 either of the expressions gives a value of Zo with-
in \% of the exact one. We thus have a solution for the characteristic impedance of
a conical strip in two expressions which between them are accurate to better than
|% throughout the whole range 6 = 0 to n/2,
(a) Z o = 2761og10(2cota/2) for0<Ji.e. Z O >188.5H (18.27)
128.6
(b) Zo = ; /nn nxx fotd>i i.e. Z o < 188.5O (18.28)

(see Fig. 18.33).


These results can be used in order to derive the Z o of a split cylinder line, as
894 Coaxial Transmission Lines and Components
shown in Fig. 18.34. In this Figure the sphere is shown tilted so that the strip trans-
mission line on the equatorial plane can be seen. The projections of the arcs where
they intersect the sphere are two arcs A'AA" and B'BB" in the tangent plane. These
subtend the same angles at the centre, Q, as do those of the strip line at 0 . This is
shown in the inset. The Z o of a line erected upon these arcs will be the same as that
of the strip line, and so the two expressions derived above together with their limi-
tations apply also to the split cylinder line with 6 defined as in Fig. 18.34.

Fig. 18.33 Characteristic impedances of a double conical sheet and a conical sheet normal to
a plane
(a) Double conical line of thin sheet
Zo = 276loglo2cot0/2 if 0 < 45° or Zo > 188.5 SI
orZ
128 6
o = (/n o
90n —
- Me\ ' f 0 > 4 5 ° or Zo < 188.5 a
Iog 10 l2cot — J
(b) Conical sheet normal to plane
Z o = 138log l o 2cot0/2 if 0 < 45° or Z o > 94.25 O
64 3
or Z = if e > 4 5 o r Z
° 7 ' or> fl\ ° « < 94-25 "

The above equations are approximate; each has its greatest error within the ranges
of 0 recommended above when 0 = 45° when the error is less than j%.

It is to be noted that with this line there is a complementary split cylinder line
of half angle (TT/2 — d) which is complementary in the sense of Babinet's principle
and which supports fields which are complementary. These bear this relationship
Coaxial Transmission Lines and Components 895
because the cross-sections of the pair can be projected to give a complementary pair
of conductors within the equitorial plane.
A number of examples are now given which illustrate the ways in which the

Fig. 18.34 Z0ofa split-cylinder transmission line


Zo = 276 log lo (2cot0/2) d 45° \.e.Z0> 188.5
128 6
" d A5° i.e. Z o < 188.5T
Iog 10 |2cot
90 -d\

to better than £%.

characteristic impedance of a new line can be derived from a known one by stereo-
graphic projection. Some of these lines are uniform and parallel and some are
conical. Thus from the expression for the Z o of the double conical sheet shown in
Fig. 18.33 we use the projection of Fig. 18.35 in order to derive theZ 0 of the uni-
form line formed by a strip between two semi-infinite planes, and we use a different
projection in Fig. 18.36 in order to obtain that of a strip parallel to a single semi-
infinite plane. Then going back to the expression for the strip between two semi-
infinite planes derived in Fig. 18.35, the projection of Fig. 18.38 gives the Z o of
the conical line formed by a conical sheet normal to a plane; in the upper part of
Fig. 18.39 this leads to the Z o of a cylinder parallel to a thin sheet, and this result is
then used with the projection of the lower part of the figure in order to give the Z o
of a cone on the edge of a semi-infinite plane sheet.
896 Coaxial Transmission Lines and Components
By this process of projecting a known line to produce another and then re-
projecting back in a different way a limitless number of conical and uniform lines
could be produced, together with expressions for their characteristic impedances.
The transformations described above will now be considered in more detail.

Fig. 18.35 Characteristic impedance of a strip between two semi-infinite planes


Z o = 138log l 0 -^ for Z o > 188.5 a

for Z o < 188.5 a

Fig. 18.36 Characteristic impedance of a strip parallel to a semi-infinite plane

Z o = 276 Iog102 forZ 0 > 188.5O


128.6
Zn — BS= forZ o < 188.5S7
7
1 +-
5

The characteristic impedance of a strip between two infinite planes can be


derived with the same \% limit using the projection of Fig. 18.35. In this case the
conical strip is placed with its axis normal to the tangent plane; as a result the two
angles marked 6/2 are equal and the two right-angled triangles of which the broken
lines form the hypotenuse are similar; hence it is seen that, where D is the diameter
of the sphere,
Coaxial Transmission Lines and Components 897

tan0/2 = -~ = — = y/d/s and cot (9/2 = y/sfd


D s/2
and therefore
Z o = 2761og lo (2cot0/2)

may be written

F ig. 18.37 Characteristic impedance of a pair o f equal parallel strips


This system of strips is the Babinet counterpart of that in Fig. 18.35. Hence the
product of the characteristic impedances in the two cases will be Z0Z'Q = TJ2/4

Zo = 276log 10 (2 for Z o > 188.512


-1

257
forZ f t < 188.5 ft

Fig. 18.38 An angular strip on a semi-infinite plane


Z o = 138 log lo (4cot0/2) f o r Z o > 188.5ft
1 28
orZ0 = / / 1 6 , / O , ,v forZ0<188.5ft
898 Coaxial Transmission Lines and Components

cylinder near edge angular strip normal


of semi-infinite sheet to a plane

cone on edge of
sheet
Fig. 18.39 Cylinder parallel to a sheet edge and a cone on the edge of a sheet
(a) For cylinder parallel to thin sheet

+1
Z o = 138log l0 2

128.6
orZ o = if Zo < 94.25 n
Iog 10 4fl
{b) For cone and sheet
for0 < 45°, Z0> 94.25
64.3
orZn = for0 > 45°, Z O < 94.25
g10
Iog 90 -e
10((2cot
Coaxial Transmission Lines and Components 899
Also, by expanding cot ((90 — 0)/2) in terms of cot 0/2
Z
o= ( 1 2 8 'L_^\\ (18-29>
logio I

may be written
7
; /
logio 2'y/s/d-I
As is noted in Fig. 18.35, the first equation applies if Z o > 188.5 and the second
if Z o < 188.5.
It is clear that if the conical strip of Fig. 18.35 were rotated clockwise so that
the point M reached the point P, then the right-hand semi-infinite plane would
move to infinity and disappear leaving the strip near a single semi-infinite plane.
This arrangement is shown in Fig. 18.36. Because of the equality of the angles
marked 0, tan0 =D/s = d/D = y/d/s; hence cot0/2 = y/s/d 4 y/l 4 s/d and the
equation
Z o = 2761og lo (2cot0/2)
can be rewritten in terms of s and d:

Z o = 2761og102 / - 4 / 1 4 - for Z o > 188.5 12 (18.31)


* d V dJ
Also, since tan ( 9 0 - 6) = s/D = D/d = y/sJdy it follows that cot ( ( 9 0 - 0)/2) =
y/dfs 4 y/\ 4 d/s and the equation
128.6

logio 12 c o t - —

may be written
128.6
Zo = J-J= r==y for Z o '< 188.5 « (18.32)
logi02

The characteristic impedance of a pair of equal parallel strips can be deduced in


a simple way from the case of Fig. 18.35. The new case is shown in Fig. 18.37. A
comparison of these two Figures shows that the two systems of strips, both those in
the spheres and of course their projections, are Babinet counterparts. Therefore the
expressions already deduced for Fig. 18.35 may be used to derive those given below
Fig. 18.37 by applying the relationship Z0ZQ = r\2/A quoted earlier.
It is clear that turning the conical strip within the sphere would widen one strip
and narrow the other. This could lead to an expression for the Z o of a line of
unequal strips.
900 Coaxial Transmission Lines and Components

The characteristic impedance of an angular strip projecting from the edge of a


semi-infinite plane is shown in Fig. 18.38. It can be seen that in this case PQM is
isosceles and right-angled. Hence s/2 equals the diameter of the sphere and
cot 0/2 = s/d. From Fig. 18.35 it then follows that the Z o of the angular strip and
the semi-infinite plane can be written down using the expressions given in that
Figure for a strip between two infinite sheets:
Z o = 1381og lo (4cot0/2) forZ o > 188.5 ohms (18.33)
or I2g 5
Zo = r /_'.„,„ , , i forZ o < 188.5 ohms (18.34)
logio 2 :

The two halves of Fig. 18.39 show that the transmission line formed by an angu-
lar strip normal to a plane is equivalent to that of a cone of equal half angle on the
edge of a semi-infinite sheet. This is done by the two projections shown. The upper
one equates the angular strip and plane to a cylinder parallel to the edge of a strip,
while the lower part shows that the cone on an edge is likewise equivalent to a
cylinder and sheet.
It is now necessary to establish some relationship between the two cylinder-
sheet lines. In fact they are both the same as can be seen:
In the upper Figure, MOQ = [Or/2) + 0]; hence 6?M = [(TT/4) + (0/2)] and
t}PB = [(TT/4) - (0/2)]; also QCP = 0.
Therefore QA = tan [(TT/4) + (0/2)], QB = tan [(ir/4) - 0/2)], OC = cot 0 =
tan[(7r/2)-0]and

,
bf (n , x ... .
tan T 4 - - -f tan - —-

which can be reduced to

tan
li"lj (1835)
v - -nn\—7rv
tan^ + - J - t a n ^ - -
In the lower Figure, BQ = tan [(JT/4) - (0/2)] and AQ = tan [(JT/4)+ (0/2)] and

(
n \

4~2
(1836)
V = I* e\ (n e^
tan^ + - j - t a n ^ - - -
Hence a'/b' = d'/b" and the two cylinder-sheet lines have the same characteristic
impedance, and therefore for the cone on sheet edge line,
Coaxial Transmission Lines and Components 901
ZQ = 1381og lo (2cot0/2) for0 < 45° or Z O > 94.25 ohms (18.37)
or
64 3
for0 > 45° or Z O < 94.25 ohms (18.38)
/90-fl
2
log10 2 cot
I
which are the same expressions as in Fig. 18.33
It will be seen en passant that it has been shown that the Zo of the line formed
by cylinder and sheet is also equal to that given by the above expressions. In order
to express Z o in terms of the dimensions of the cylinder-sheet line it is convenient
to introduce st and s2 as in Fig. 18.39; then it can be shown that, for the cylinder
and sheet,

= 1381o gl0 2 i f Z 0 > 94.25 ohms (18.39)

or
10R ft
ZQ = 1_ i f Z 0 < 94.25 ohms (18.40)
1
log io 4 -

The upper expression can be rearranged in terms of the diameter d and axial
spacing from the edge, y, so as to give

= 1381og 10 2|— + / — | - 1 ) ifZ 0 >94.25 ohms (18.41)


Kd V \ dt
This will be recognised as the expression for a cylinder parallel to a plane with
the addition of a 2 in front of the bracket. This difference adds an extra 1381og102
(= 41.6)12 to the Z o of the cylinder-edge line compared with that of the cylinder
and plane line of the same ratio of cylinder diameter to gap. This difference is illus-
trated in Fig. 18.40, where the curve of Z o for the cylinder and edge runs a constant
41.612 above that for a cylinder and plane. There is likewise the difference of a fac-
tor of 2 in the log terms of the expressions for the Z o of a cone on an edge and a
cone normal to a plane, introducing, therefore, the same difference of 41.612. In
each case, replacing a plane by an edge adds 41.612 to Z o . This change is met in
practice when an antenna designed to work on a plane is sited on the edge of an
aerofoil section.
It should be noted that, in all the cases where pairs of expressions are given to be
used above or below a limiting angle or impedance, the greatest error is at this
limit where it is less than \%, and that it falls to negligible levels at the lower and
upper ends of the range of Z o . The transformations used are all exact; this small
error arises from the approximations used in determining the expressions in Fig.
18.33.
902 Coaxial Transmission Lines and Components

18.6 Calculation of characteristic impedances

18.6.1 Constants in the expressions for Zo


In preparing the expressions for Z o used in this Chapter it was decided to continue
the well-established custom of engineers to favour common rather than Naperian
logarithms and to combine various products of n and loge10 into numerical
constants which are rounded off. In Fig. 18.41 these constants are tabulated
together with their functions of n and/or loge10. The last column shows their
values rounded to seven figures.

300

200

o
INI

100

cylinder and edge cylinder and plane


(for ZQsee upper curve.) (forZ Q see lower curve.)

2 U 6 8 10 12 tt 16 18 20

gap from edge or plane / cylinder-diameter

Fig. 18.40 Comparison of the characteristic impedances of a cylinder near an edge and near a
plane
For cylinder and edge

For cylinder and plane


Zo = SOcosh" 1 ^ = 138log l0 (?f

18.6.2 Range of validity for approximate formulae for Z o


There are many published tables of collected expressions for the Z o of different
transmission lines. It is a general defect of these that they quote formulae without
Coaxial Transmission Lines and Components 903
any indication of the range of spacing to cross-section for which they are accurate.
Thus for the formulae for most open-wire lines there is the caveat that the wire
diameter should be very much smaller than the spacing of the conductors, but no
indication of how inaccurate the calculated values would be for any given close
spacing. It is not easy to trace the original derivations of these formulae in order
that their limits of accuracy can be assessed. However, some rough working rules
can be deduced.

Approx. value Function of n or logg 10 Value to 7 figures


377 120?r 376.9911
276 120loge10 276.3102
188.5 60?r 188.4956
173 75log e 10 172.6939
138 60log c 10 138.1551
128.6 30?r2/loge10 128.5894
94.25 30?r 94.24778
69 30log e 10 69.07755
64.3 157r2/ioge10 64.29472
41.6 60(loge 10) (Iogi02) 41.58883
1.07 201 1.071773

Fig. 18.41 Values of the numerical constants in the expressions for Zo

Consider the case of the two-wire balanced line shown in Fig. 18.11 for which
the characteristic impedance is plotted in Fig. 18.12. When the two conductors are
far apart the current is evenly distributed around the periphery of their circular
cross-sections. In these conditions the simple formula

Zo = 2761o g l 0 ^j (18.42)
a
applies and we have the right-hand part of the curve of Fig. 18.12.
As the conductors come closer together so that diameter and spacing become
comparable, the currents concentrate upon the parts of the wires which are close
together. This reduces Z o below that which would apply if the distribution were
uniform. For such close spacings it is essential to use the exact formula

Z o = 120 cosh"1— (18.43)


d
and it is this which has been plotted in the left-hand part of Fig. 18.12, where the
rapid drop in Z o as the spacing decreases is the result of the concentration of cur-
rent. It can be demonstrated that, when the axial spacing is 3.7 times the conductor
diameter, Z o calculated by the approximate formula will be 1% in error and the per-
centage error rises very rapidly with decrease in this ratio.
904 Coaxial Transmission Lines and Components
Consider now the case of the 3-wire line of Fig. 18.16a. When the wires approach
each other the current distribution on the outer pair of wires will suffer a concen-
tration similar to that in a balanced pair, but the inner wire will have a current
maximum on each side and so the overall change will be less. Hence one could
apply the two-wire criterion that the spacing should be at least 3.7 times the wire
diameter for less than 1% error with some small factor of safety. Similar reasoning
can be developed in order to show that the limits of accuracy are much the same
for the other types of line, the factors of safety being somewhat greater.
Hence for open-wire lines in general, it can be said that the approximate expres-
sions for ZQ will be accurate to 1% provided that the ratio of axial spacing to wire
diameter is greater than 3.7; below this spacing the true values will fall away rapidly
as does the curve of Fig. 18.12.
The above result can be extended to the cases where a conductor is parallel to a
conducting plane. In this case one applies the criterion to the spacing of the
conductor from its image in the plane and modifies the rule of thumb to require that
the spacing of the conductor from the plane should be greater than 1.85 times the
conductor diameter for 1% accuracy.
In the case of the wire between parallel planes of Fig. 18.22a a correction term
for close spacing is given. It is easy to show that the neglect of this gives a 1% error
at a spacing of 1.8 times the conductor diameter. This is a pleasing confirmation of
the above rule of thumb.

18.6.3 Methods of calculating or estimating the Z o of other lines


If the cross-section of a line does not conform to a standard shape there are various
ways in which its Zo can be calculated using an exact method, or at least estimated
with reasonable accuracy for most antenna design work.
In the case where single circular conductors are replaced by stranded conductors
or those of strip, square or angled cross-section, the correction to Z o is simple and is
given in Figs. 18.20 and 18.21.
Methods of introducing conducting surfaces in the place of equipotential sur-
faces in order to obtain a new conductor geometry have been demonstrated in
earlier Sections of this Chapter; in this way the characteristic impedances of the
lines of Figs. 18.16c and \S.\6d have been derived from that known for Fig. 18.16&.
In the Section dealing with conical transmission lines many examples are given
of the method of developing a new line from a known one by stereographic projec-
tion and determining the Z o of the new one from that already known. This projec-
tion is one class of conformal transformation which can be used in order to deter-
mine ZQ. The basic theory of the calculation of Z o by means of conformal
transformation has been discussed by Anderson;23 its application is illustrated in a
number of sources given in the bibliography (Magus and Oberhettinger,25 Morton,26
J. J. Thompson28).
Many of the derivations of the Z o for the standard forms of transmission line
were originally made in order to derive the capacitance per unit length associated
with conductor systems, and these will be found in textbooks on electrostatics.
Coaxial Transmission Lines and Components 905
Where exact methods are not possible, various approximate means of estimating
can be used. For very complex cross-sections field plotting can give results where
other methods fail. The procedure for this method is discussed by Willoughby.29
With computer programs available this rather tedious method would not be used as
a routine process although it provides a good way of becoming familiar with the
basic principles. This paper describes field plotting by eye together with electrolytic
tank and other methods.
It is often useful to remember that the characteristic impedance of a wave travel-
ling on a transmission line is equal to the resistance between the conductors if they
were joined by a resistive film of 377 O/square. In the simple case of Fig. 18.24&
with the condition d/b < 1, such a consideration leads immediately to the expres-
sion Z o =* 377d/b. In the case of a uniform line the wavefront will be plane and so
will be the postulated resistive film; however, for a conical line supporting a
spherical wavefront, the resistive film must be imagined as lying within the wave-
front (i.e. on a spherical surface). It is this change from plane to spherical which is
allowed for by the stereographic projection from a line having a conical set of con-
ductors to that of its uniform counterpart.

Fig. 18.42 Square-section conductor and equivalent circle


Diameter of circle = 1.18 times side of square

Consideration of the resistive film is of use when assessing the effect of the de-
formation of one line in order to form another. For instance, take the square-
section conductor shown in Fig. 18.42. As noted in Fig. 18.21, this is equivalent
for the purpose of calculating Z o to a circle of diameter 1.18 times the side of the
square. Such a circle is shown. If one considers this in terms of the hypothetical
resistive film, it is seen that in going from the square to the circle one removes the
corner pieces (which are cross-hatched) and replaces them by the segments (which
are shaded). It will be seen that the two arrangements are comparable, particularly
when it is borne in mind that, while the corners are slightly smaller than the seg-
ments, there would be a concentration of current on them, or one could say that
being nearest to the outer conductor the film resistance between them and the
outer conductor was least.
906 Coaxial Transmission Lines and Components
18.7 Layout of cable runs for minimum interference

Earlier in this Chapter the quality of cable screens was discussed and the concept of
the 'transfer impedance' ZT was introduced. Fig. 18.43 shows a cable acting as a
source of interference. If the current flowing in this is /, then the signal voltage on
the outside developed by leakage through the screen is:
VT = ZTI volt/unit length (18.44)

Fig. 18.43 Leakage of signal from one cable into another


(a) Interfering cable
For signal current /, signal voltage on outside of screen = V^ — Z^l volt/unit
length
(b) Cable interfered with
Signal voltage on inside of screen = Vjp~ Z j - / ' volt/unit length

This voltage will induce a current / ' on the outside of the screen of a second cable,
and thereby develop a voltage V*T — Z^l' per unit length on the inside of this cable.
The reduction of ZT and Z'T by the improvement of the cable screens has been dis-
cussed; however, it is just as important to minimise the coupling between the out-
sides of the cables so that / ' shall be as small as possible for a given VT. This is the
next matter to be discussed.

18.7.1 Coupling between coaxial cables


We begin by considering the coupling between two coaxial cables running parallel
to each other and to a conducting plane as shown in Fig. 18.44. It is convenient to
build upon the analysis published by Mohr.3 In this he considered the coupling
between two parallel cables such as AB and CD of Fig. 18.44. In the first paper he
assumed that the working frequency was low enough for skin effect to be neglec-
ted, but in the second the analysis is extended to cover the frequency range up to
about 50 MHz. This upper limit in frequency is set by the increasing effects of the
resonances in the external circuits as the cable lengths employed become appre-
ciable fractions of a wavelength. This limitation was of importance to Mohr, who
required a quantitative theory capable of experimental verification. However, the
resonances would not vitiate any conclusions about the relative merits of different
layouts at higher frequencies.
Coaxial Transmission Lines and Components 907
Mohr's approach is to consider the arrangement of Fig. 18.44 and to split it up
into three sub-circuits:
(a) A loop of inductance V per unit length between AB and the plane
(b) A loop of inductance L" per unit length between CD and the plane
(c) A mutual impedance of M per unit length linking the two loops;
It is assumed that M is small, otherwise the circuits could not be treated separ-
ately.

o oT
Fig. 18.44 Two parallel cables spaced from a conducting plane

Mohr gives curves from which L\ L" and M can be read given the diameter of the
cable screen d, the height of the cable axis from the screen h, and D the spacing of
the cable axes. He derives an expression for Vd, the interfering voltage in CD, which
is given here, with some simplification valid at radio frequencies;

Vd = coi- (18.45)
L'L"
where / = current in AB, Z'T and Z? = transfer impedances of AB and CD, Rc and
Rd = terminating impedances of CD.
This expression shows, as one would expect, that the interfering voltage is pro-
portional to the expression
M/L'L"
which is a parameter dependent upon the local geometry, i.e. the installation layout.
908 Coaxial Transmission Lines and Components
In order to display the variation of M/L'L", Fig. 18.45 has been prepared. The
solid curve refers to a pair of Uniradio 43 cables (equivalent to RG-58/U) rest-
ing on a conducting plane; the value of h = 5 mm represents the case where the
cable screen is separated from the plane by the thickness of the PVC jacket. (For
simplicity, the fact that the dielectric constant of PVC is greater than unity is
ignored). At the left-hand end of the curve the pair of cables are close with their

10-

10-

0-1- forUR 43 d = 3-3mm

h = 5mm

001
25 50
D.the spacing of the cable axes, mm

Fig. 18.45 Variation of M/L 2 with cable spacing


For UR43, d= 3.3 mm

jackets in contact. It is seen that in this position M/LfLff* is great, but that as the
cables are separated this parameter falls rapidly, and when their spacing is 50 mm it
has been decreased by a factor of 80. This means that the interfering voltage has
decreased by the same factor and the power coupled decreased by 18 dB. This
demonstrates the importance of cable separation in determining levels of interfer-
ence.

* As V = L" in this case we use the parameter M/L2 later in this Section and in Fig. 18.45.
Coaxial Transmission Lines and Components 909
2
In the same figure the dashed curve shows the variation of M/L with spacing for
a pair of cables 25 mm from the sheet. It is evident that the decrease in M/L2 is less
marked than when the cables are close to the plane. The dotted curve shows that at
a distance of 500 mm away from the plane there is still less reduction in coupling
on moving the cables apart. Comparing the three curves, it is seen that the marked
decrease in M/L2 as the cables are separated only occurs when the cables are very
close to the sheet, a closing in from 500 mm to 25 mm bringing but a small
improvement compared with that obtained in the last part of the approach to the
plane. It is clear that the effects of the dielectric constant of the PVC outer layer
would be to give an even greater benefit from close contact of the cable and the
plane that was calculated when the data for Fig. 18.45 were prepared.
From these considerations one can deduce two rules for the installation of cable
runs near to conducting sheets:
(i) The cables should be held as close as possible to the sheets
(ii) Cables which might interfere with each other should be held as far apart as pos-
sible.
The above discussion does not take into account the effects of resonance when
the sections of the cable between earthing points are an appreciable fraction of a
wavelength long. The effects of resonance are difficult to assess because the
Q-factor of the circuits between the cable screen and the earth sheet are not easy to
define. However, general principles for guidance are clear:
(i) Because the effect of resonance is only to apply a multiplying factor, the coupl-
ing between the circuits will still be reducible by the choice of cables with low
values of ZTi and by good layout which minimises M/L2.
(ii) At the highest frequencies the losses in PVC sheaths will tend to damp the
resonances, and this damping will be increased the closer the cable is held to the
conducting screen because of the concentration of the field within the PVC.
After considering the coupling between two cables near to a conducting plane it
is useful to study that between those installed in a metal cylinder such as the fuse-
lage of an aircraft. In Fig. 18.46^ is shown a cable along the axis of the cylinder.
For any TEM wave guided by the cable within the cylinder the equipotential sur-
faces will be as represented by the broken circles. If the cable is moved to the left as
in Fig. 18.46& then the field remains TEM but the equipotential surfaces will be
squeezed together. They will remain circular and will form a family for each of
which
OAxOB = OP2
It will be noted that the displacement of the inner conductor increases the volt-
age gradient on one side but decreases it on the other. An extreme case is shown in
Fig. 18.46c, where a Uniradio 43 cable is held with its PVC touching the wall
of the cylinder; the centre of this diagram is shown in Fig. 18.46d It is seen that
within the region about A A' the conductors can be taken to be parallel planes and
therefore the voltage gradient may be taken as uniform across AA\ Corresponding
910 Coaxial Transmission Lines and Components
to the various points A n on the line AA' (see Fig. 18.46Z?) are equipotential points
B n . As the potentials of the An points are known because of the uniform gradient,
the potentials of points B n can also be found and the voltage variation across the
diameter of the cylinder be determined.

pvc sheath

Fig. 18.46 Field around a cable in a fuselage

Using this approach, the variation of voltage gradient across the diameter of a
fuselage 2.44m wide has been calculated for two cases: where the Uniradio 43
cable touches the wall at one end of the diameter and secondly where the axis of
the cable is 50mm away from the wall. The data are plotted in Fig. 18.47 where
the voltage gradient in volts/metre is shown for an assumed potential difference of
1.0 V between the cable and fuselage. Curve a is for the cable touching the fuselage
while curve b is for one spaced 50 mm away. For the close-spaced cable the greatest
Coaxial Transmission Lines and Components 911
voltage gradient between screen and fuselage is 245 V/rn, much greater than the
maximum of 30 V/m for the other one, but curve a falls away so rapidly that at
only 50 mm away from the wall the gradient is only 2 V/m and it falls below

100 1

0001
0-5 10 15 20 2-5
distance along fuselage diameter, m

Fig. 18.47 Variation of voltage gradient across the diameter of a fuselage


(a) Uniradio 43 touching wall
(b) Cable axis 50 mm from wall

0.001 V/m at distances beyond 2.0m along the diameter. At the far end of the
diameter the voltage gradient for cuve a is only l/25th of that for curve b.
Consider a second cable at the other end of the diameter; the potential difference
912 Coaxial Transmission Lines and Components

between it and the wall would be the voltage gradient at the wall multiplied by the
distance of the axis from the wall. Thus, if a pair of cables were put one at each end
of the diameter with sheaths touching the wall and then both were moved to have
their axes 50 mm away from the wall, the voltage gradient at the second cable
would rise by a factor of 25; and because the axial separation from the wall has
increased by 10 (i.e. from 5 mm to 50 mm) the voltage induced on the cable would
increase by a factor of 250. The power picked up in the second cable would thus
increase by (250)2 or 48 dB.*
Once again one sees the importance of keeping cables close to the conducting
screen, this time for the two reasons
(a) In order to reduce the voltage gradient in the main volume of the fuselage
(b) In order to minimise the mutual coupling between cables
In some cases an installation specification sets a maximum limit on the voltage
gradient within the fuselage due to signal leakage. The requirements of the specifi-
cation should be worded so that the gradient in the immediate vicinity of the cable
is not limited; otherwise one might exclude the possibility of mounting a cable
close to the screen as for curve a in favour of an installation such as that for curve
b. This choice would, in the case considered above, reduce the maximum gradient
existing at the cable surface by a factor of about 6 at the expense of increasing the
gradient in about 99.99% of the rest of the volume of the fuselage by a factor of
about 20.

18.7.2 Mutual screening of vulnerable screened circuits


By placing a second cable adjacent to a cable carrying a signal which must be
screened from external interference, the signals picked up in each of the two cables
are less than those which would be picked up by either in the absence of the other.
Thus there is a mutual screening — as it were, they help each other.
This mutual screening comes about through the mechanism illustrated in Fig.
18.48. In (a) is shown a conductor A near to a conducting sheet; AA' is a line of
force cutting an equipotential circle PQ at B. One could place a thin perfectly con-
ducting sheet in the surface PQ without disturbing the field, and likewise could
place an infinitely thin perfectly conducting wire at B without causing any dis-
turbance. Now if this wire were to swell to the size of the conductor B as in Fig.
18.48&, the tubes of magnetic induction between B and A' would be compressed.
This extra pressure would be relieved by a transfer of charge from the A' side to the
A side of B and an associated decrease in the induction on the A' side and increase
on the A side. As a result, the mutual coupling between A and B would decrease as
the diameter of B increased. Because of the pinning back of the tubes of induction
between B and A' and remembering the internal pressure of JJLH2/Sir of these tubes

This assumes that the presence of the cable does not distort the field which would exist in its
absence. This is not exactly so and the assumption tends to reduce the ratio of induced voltages
so that the figure of 10 deduced is an underestimate.
Coaxial Transmission Lines and Components 913
it is clear that the equipotential surface which was circular in Fig. ISASa will be
further distorted by a swing nearer towards A in the region of Q.
If now a conductor C is introduced as in Fig. 18.48e, so that B and C are sym-
metrically placed with respect to A and the plane, then both must be at the same
potential and PQ must be pinned back by C as well as by B. Again the increased
pressure between the tubes of induction on the plane side of the conductors results
in peripheral currents on B and C which allow tubes of force to move toward A and
further reduce the mutual inductance between A and B. It is clear that the mutual
inductance (or coupling) between A and B has been decreased just by the presence
of C — and likewise the coupling between A and C is less because of the presence of
B.

F I G . 18.48 Mutual screening of adjacent cables

This is what is meant by 'mutual screening' — that by their presence adjacent


cables can reduce the interfering signals induced by an intruding field. It is not easy
to estimate the likely benefit of mutual screening, but it is likely in general to be a
few decibels.
Gathering our conclusions about the layout of cable runs, one concludes:
(i) All cable runs whether at high power and acting as potential sources of interfer-
ence, or at low power and vulnerable to interference, should be as close to the
earthed conducting screen as possible.
(ii) All cables which are potential sources of interference should be as far as possible
from those cables which must be protected.
(iii) The cables to be protected should be run side by side as close to an earthed
conducting plane as possible in order to benefit from mutual screening.
Before leaving the question of interference within metallic structures it should
be noted that, as the frequency of operation increases and the cross-section of the
structure becomes comparable to a half wavelength, there is the possibility of parts
of the structure acting as waveguide resonators and so enhancing the local field
strength of any interfering signal. Here two comments are made:
914 Coaxial Transmission Lines and Components

(i) In any waveguide mode of resonance, the field strength near to a metallic boun-
dary will be low — yet again one sees the advantage of running cables close to a
metal screen.
(ii) The lower cut-off frequency for any structure such as an aircraft fuselage will be
at its maximum when all metallic items such as the frames of seats and the metal
racks for non-radio items such as safety equipment are as well bonded electrically
to the fuselage as are the units holding the radio and electronic equipment. Each
unit well bonded reduces the volume which is free to resonate.
Because this Chapter is about transmission lines, interference has been discussed
as if it began and ended with cable runs. In fact, a very large number of interference
problems arise through leakage from the units of the radio equipment rather than
from the cables which join them. It is often thought that, if touching the outside of
a cable changes the level of interference in a given installation, then a signal has
leaked out through the cable screen. In fact, all that it means is that a signal is flow-
ing on the outside of the cable — in many cases it gets there through defects in the
screening of the equipment boxes and not through a deficient cable braid.
The screening of plugs and sockets is discussed later.

18.8 Junctions in coaxial line systems

18.8.1 Discon tinuities at junctions


When it is necessary to join together two dissimilar lines in series various problems
must be solved if a significant junction discontinuity is to be avoided. Consider first
the case where the lines are of the same characteristic impedance but differ in size.
Such a pair of lines are shown in Fig. 18.49#. If these are just butted together as in
Fig. 18.49&, in which the outer conductors are jointed by a ring which lies in the
same plane AB as does the junction of the two inner conductors, then two capaci-
tive discontinuities are introduced, the first by the capacitance between the step
between the outer conductors and the larger inner conductor and the second by the
capacitance between the step on the inner conductor and the smaller outer conduc-
tor. This parallel capacitance can be compensated by extending the smaller inner
conductor into the larger outer conductor by the distance 5 shown in Fig. 18.49c.
When this is done the ratio of conductor diameters in the section of line ABCD is
larger than that on either side of the junction, and for this reason has an increased
inductance per unit length. The displacement 6 is chosen so that this increase in
inductance compensates for the extra capacitance of the discontinuities, and
the ratio of inductance to capacitance (i.e. Z\) is kept constant through the junc-
tion.
Values for the capacitance associated with step discontinuities were derived by
Whinnery, Jamieson and Robins30 in 1944 and later by other authors including
Miles.31 The method of correcting by displacing the junctions, as in Fig. 18.49c, was
due to Meinke32 who published data in 1943. Further measured data have been
published by Krauss.33
Coaxial Transmission Lines and Components 915

The pairs of lines can be joined by a coned section such as that shown in Fig.
18.49d. The cones are proportioned so as to join the conductors within the planes
AB and CD; as a result, their apices are slightly displaced longitudinally. For this

B
A

Fig. 18.49 Junctions between coaxial lines of the same Zo but different sizes
(d) For 1% match keep half angle of outer cone < 20°
For 4% match keep half angle of outer cone < 40°

reason the section differs from the conical line shown in Fig. 18.27fl;but if this
slight difference is neglected and if the half angle of the outer cone is kept below
20°, one can equate tan0/2 to 0/2 to an accuracy of 1% and so obtain a match to
within 1%. It will be noted that even if the Z o of the conical line were equal to that
of the adjacent lines, there would still be small discontinuities at the ends of the
cones, where the spherical waves on the conical section meet the plane waves on the
uniform sections. The slight displacement of the apices of the cones gives an excess
916 Coaxial Transmission Lines and Components

capacitance at the ends of the conical section which tends to correct for the change
from plane to spherical wave fronts.

18.8.2 Impedance-matching sections


In the case where the lines to be joined differ in Z o , an impedance matching section
is required. Of these the most usual is the quarter-wave transformer of Fig. 18.50a.
This transformer is a section of line whose Z o is the geometric mean of these two
lines, and it matches at a given frequency when it is X/4 long. In the lower part of
the Figure is shown the impedance changes along AB plotted on the Cartesian form
of circle diagram;* this shows how the impedance across a transmission line varies
along the line. Suppose that the transmission line to the left of A is perfectly
matched. Starting from A, where the impedance is Zo, the impedance at inter-
mediate points lies upon a semicircle ending at B where the impedance is Zo, the
required match for the right-hand line. Now if the frequency is reduced, the length
of the transformer will be less than X/4 and B will not reach the axis; likewise, for a
higher frequency the transformer will be too long and B will pass the axis taking the
impedance to B', again resulting in imperfect matching.

z'o zo z'b z'o zoi z6


(a) (b)

1 I
X
xR B' tx &'
— / 0 A i
i i

;
\ ,

Impedances normalised Impedances


Impedances
toZ0 normalised t o Z 01
normalised to Z o ,
Fig. 18.50 Single and double quarter-wave transformers
(a) Match if Zrto = sf^J;
JZ'nZ"
(b) Match if Z01Z02 =Z'0Zi'
Broadband if Z 0 1 = [(Z;) 3 Z;'] 1 / 4 ; Z 0 2 = [Z^Z;') 3 ] 1 ' 4

This effect can be reduced by having two sections AB and CD each X/4 long in
series. In this case, referring to the lower part of Fig. 18.50&, the impedance trans-
formation takes place in two stages; firstly following a semicircular arc AB in the
diagram of impedance normalised to ZOi, and then following a second arc BC in the
diagram normalised to Z02- If the frequency is increased so that the lengths AB, BC
are more than X/4 long the first length takes the impedance into the inductive part
* The Cartesian form of the transmission-line circle diagram has been discussed by a number of
authors including Willis Jackson.34
Coaxial Transmission Lines and Components 917

of the diagram to B', then the second half will start in the inductive region at B',
pass some way along its length before the axis is reached and then end up near to C.
Thus the error introduced in AB is compensated in BC. The optimum ratio of the
impedances ZQ and Zo" to be chosen for the X/4 lengths depends upon the desired
frequency bandwidth and the impedance transformation ratio ZQ/ZQ required. It
has been shown by R. A, Smith14 that, when simplifying assumptions are made, the
greatest bandwidth is given when the transformation ratio of each X/4 length is the
same, that is:
Zj = [(Zf0)3Z£]m and 2% = [Zj(Z^) 3 ] 1/4 (18.46)
In general the improvement on this solution which can be expected from an exact
solution is small, and also likely to be obscured by the inevitable junction discon-
tinuities between the lines of differing cross-section. It may be noted that the above
solution has advantages when high-power working is required because the VSWR in
the X/4 sections is at a minimum.
Where a transformation from one Z o to another is required to operate over a
wide frequency it is usual to use a tapered section. This is made as long as con-
venient - say a wavelength at the lower end of the frequency band. With this
arrangement there is a small discontinuity associated with each short length of the
taper. These are all reflected back to the input with progressively lagging phase, and
so are subject to destructive interference - hence the resultant VSWR at the input
is small. This principle of securing broadbanding by spreading a discontinuity pro-
gressively along a length of line is of wide application. In the special case where the
change in Z o per unit length is constant, one has the exponentially tapered line.
This has attracted considerable attention and been discussed first by Wheeler35 and
then by others including Christiansen,36 Niutta37 and Rohrman.38
The construction of a truly exponential line may present difficulties in produc-
tion and not offer any great advantage over more simple forms of taper such as
those discussed by Gurley,39 by Klopfenstein40 and by Gent and Wallis.41

18.9 Balancing devices or baluns

The value of an impedance is defined when the current produced in it by a differen-


tial voltage between its ends is known. However, this value is not all that it is neces-
sary to know if an impedance is to be fed correctly by a transmission line. Because
no impedance can be completely isolated from objects at earth potential, it is
important to know what coupling there is between the two ends of the impedance
and earth.
These stray couplings to earth can be represented by the equivalent circuit of
Fig. 18.51. There are two special cases which must be considered in detail: the first
is when Z 2 = Z 3 in both magnitude and in phase. In this case the impedance AB is
said to be 'balanced' and the two potential differences between A to earth and B to
earth are equal in magnitude and opposite in phase.
918 Coaxial Transmission Lines and Components

The second special case is when either Z 2 or Z 3 is zero. In this case one end of
AB is at earth potential and the impedance is often described as 'unbalanced'. This
term is unfortunate because its use implies that all impedances are either 'balanced'
or 'unbalanced', while, in fact, these are but two special cases among an infinity of
arrangements of the equivalent circuit of Fig. 18.51.

Fig. 18.51 Equivalent circuit of an impedance AB with couplings to earth

If one considers TEM-mode transmission lines they mostly fall into two classes:
those with a balanced and those having an unbalanced characteristic impedance.
The most common forms of the balanced lines are the open-wire pairs and the
screened twin cable. The unbalanced lines are usually but not always of coaxial
form. It would be possible to have other lines such as the unbalanced line, which
would result if an open-wire pair were made of one thick wire and one thin one;
however, such a line would not be generally useful.
In order to simplify their connection to the balanced or unbalanced transmis-
sion lines it is usual to design antennas so that their input impedances are either
balanced or unbalanced. This is achieved by ensuring the symmetry of the antenna
where a balanced input is required and by choosing a feed point at, an earth
potential point where an unbalanced input is needed.

Fig. 18.52 Examples of the types of antenna feed


(a) \ / 2 dipole centre fed
(b) Unipole fed at base
(c) X/2 dipole with offset feed

These points are illustrated in Fig. 18.52 in which (a) shows the most simple of
the antennas with a balanced input — the centre fed A/2 dipole; (b) shows a unipole
where the feed point is shown at the earthy end in order to present an unbalanced
input; (c) shows a A/2 dipole fed partway between centre and end; its unbalanced
impedance cannot be fed satisfactorily with either a balanced or an unbalanced line.
It sometimes happens that it is necessary to feed an antenna with a balanced
Coaxial Transmission Lines and Components 919

input from an unbalanced line such as a coaxial cable, also, although, less fre-
quently, an unbalanced antenna from a balanced line. Such connections can be
made without ill effect provided balancing units or baluns are used. Without such a
unit serious troubles could arise. Thus, for a coaxial cable connected directly to a
dipole, the antenna would provide a coupling between the inside and outside sur-
faces of the outer conductor. In the transmitting mode the currents induced upon
the outside of the cable would radiate in unwanted directions; also in the receiving
mode any interference signals induced upon the outside of the screen would be
coupled inside and fed to the receiver input.
The most generally used class of balun is that based upon a shorted X/4 of line
placed as a sleeve around a coaxial line. Such an arrangement is shown in (a) and
(b) of Fig. 18.53. The X/4 sleeve is open-circuited at AB and short-circuited at the
other end. For this reason there is a high impedance (which would be infinite were
it not for losses) between B and the outside of the sleeve. This being so, a balanced
impedance can be connected to BC without any coupling to the outside of the
sleeve or cable.
A simplification of the sleeve balun occurs when the sleeve is replaced by an
open stub as in Fig. 18.53c. It is clear that rotation of this about the axis through
C would generate a balun of the sleeve form. While the balanced load could be con-
nected across BC as with the balun of Fig. 18.53a, it is usual to connect A to C by a
short link and then connect the load across AB. This arrangement has the advantage
of greater bandwidth.
There are a large number of variations of the X/4 sleeve balun and information
about many of these is given by Jasik.16 In some designs of antenna, advantage is
taken of a X/4 sleeve which exists as part of the radiators in order to provide balanc-
ing. An instance of this is shown in fig. 18.53d. In this arrangement the coaxial feed
is led along the axis of one arm of the dipole; this arm itself forms a X/4 sleeve
around the coaxial line and so acts as a balun.
It will be noticed that the units so far described have not effected any imped-
ance transformation; for instance, the system of Fig. 18.53& is matched by joining
the points BC by an impedance equal to theZ 0 of the coaxial line. There are other
balancing units which introduce an impedance transformation so as to give match-
ing loads of 2Z0 or 4Z0. The 'rat race' described later in this Chapter can be connec-
ted as a balancing unit with a 2Z0 match. The arrangement shown in Fig. 18.54
gives a match with 4Z0.
The core of the 4:1 transformer is a half wavelength of line AB; because of its
length the voltages at A and B will be in antiphase. Let there be a terminating resist-
ance of 2Z0 at B; this will present an impedance of 2Z0 at A. If a second load of
2Z0 is added at A as shown in Fig. 18.54#, the resultant will be a termination of Z o
presented to the line A A', which will thus be matched.
If the line is now bent into a circle and A and B brought together as in Fig.
18.54&, then, because the voltages across AC and BC are equal and in antiphase, the
pair of loads AB form a balanced load of 4Z0. Hence a balanced transmission line of
characteristic impedance 4Z0 could be connected at AB without disturbing the
920 Coaxial Transmission Lines and Components

impedance match on AA' or exciting stray currents on the outside of the coaxial
part of the system.
Because of the 4:1 step up in impedance this balun is of use where a coaxial line
must be connected to an open-wire line; for instance, a 75 Q coaxial cable can be
connected to a 300 £2 open-wire line.

A B C

T
B C

Fig. 18.53 Quarter-wave balancing units

Where it is desired to have a wideband balun the devices so far mentioned cannot
be used. In these circumstances a tapered balun is sometimes of help. This has the
form shown in Fig. 18.55. Here the outer conductor of a coaxial line is sliced
Coaxial Transmission Lines and Components 921
diagonally over a length AB which is long compared with a half wavelength. At A,
the inner conductor and the pointed end of the outer conductor are connected to
an open-wire line. One can look upon its operation as a process of guiding whereby
the wavefront within the coaxial line at B spreads out as it approaches A so as to fill
the region about the balanced line; or for a wave coming from A to B, the remnant
of the outer conductor guides the wave, wrapping it up into a bundle small enough
to enter the coaxial line. The taper can also act as an impedance transformer, and
there is no special limit on the transformation ratio which this balun can be
designed to give.

2Z 0 2Z 0

B,

characteristic
impedance of
line = Zo

For matching
Resistor AC = 2Z0
Resistor BC = 2Z0
Hence terminating resistance AB = 4Z0
Fig. 18.54 4:1 transformer and tapered baluns

Another type of broadband element is the compensated balun devised first by


Marchand42 and developed by others including Cloete43'44 who has given a biblio-
graphy.44
These baluns can also act as impedance transformers. The design curves given
by Cloete44 have been tested experimentally and shown to give a maximum
922 Coaxfa/ Transmission Lines and Components

reflection coefficient at the coaxial input of —13 dB for a balun designed to match
from 50 £1 coaxial to 10012 balanced over the band 0.2 to 2.0 GHz. The data given
do not include the performance as a balun; only the input reflection is quoted (i.e.
only its performance as an impedance transformer is given). It should be noted that
at some frequencies in the 10:1 band resonances in the structure could make the
quality of the balancing action very sensitive to the mechanical tolerances achiev-
able in production.

Fig. 18.55 The tapered balun

18.10 Power dividers

18.10.1 The'rat race'


The rat race is a circular loop of transmission line of 3X/2 in circumference. It is
shown in Fig. 18.56 where T connections into the loop are shown at A, B, C and D,
points which are spaced as in the diagram.

Fig. 18.56 The rat race

The rat race has several uses. First, it can be used as a power divider in order to
give two signals equal in amplitude and phase. In this case, consider power fed in at
C, this will split into two equal parts; one will travel clockwise towards B and the
other counter-clockwise towards D. The signal reaching B by the clockwise path
will be delayed by TT/2 with respect to the phase at C. The signal passing counter-
clockwise will reach B by a path of 5X/4 having a phase delay of 57r/2, and so be in
phase with the signal arriving by the shorter path. In a like manner both paths will
provide adding contributions at D, their sum like that at B being delayed by TT/2
Coaxial Transmission Lines and Components 923
with respect to the phase at C; hence the two outputs at B and D are in phase. It is
clear that the two paths from C to A will give phase changes differing by TT/2, and
that consequently they will have a zero resultant at this point; for this reason, no
signal will leave the loop except at B and C. Thus the power input at C will divide
equally in amplitude and phase between the outlet ports B and D.
The second use of the rat race is to form the sum and difference of a pair of sig-
nals. If two signals are fed into the loop, one at K and the other at D, it is easy to
show by arguments analogous to those of the previous paragraph, that their sum
emerges from C and their difference from A.
In the design of some variable phase shifters, the two previous uses of the rat
race are pombined. First a signal is fed into C; this splits equally between B and D.
Two short-circuited lines of variable length are connected at B and D, the lines
being ganged so that one is always A/4 longer than the other. These short-circuited
lines reflect the signals back to B and D with two different phase changes, the sig-
nal from the line which is A/4 longer being delayed in phase by n radians. Because
of this difference in phase, the two signals will appear at the outlet A. Variation in
the lengths of the ganged lines will be apparent as phase changes in the signal at A.
If a signal is fed into the loop at A it will divide equally between outputs at B
and D, the two signals being in antiphase. This being so, the voltages at B and D are
symmetrical with respect to earth, and therefore a balanced impedance could be
connected across these points. It will be noted that, for impedance matching, the
balanced impedance should be 2Z0, e a c h half matching one of the two outputs
across which the connection is made. Used in this way, the rat race gives a balancing
unit or balun with an impedance transformation by a factor of two.
Impedance matching within the rat race is achieved when the characteristic im-
pedance of the loop is equal to \/2Z0> where Z o is the characteristic impedance of
the input and output lines at A, B, C and D. If these lines have Z o = 50 £1, that of
the loop should be 71 £2 although the use of a standard 75 £2 line would not result
in a significant error.
The rat race is mostly formed from waveguide or strip line for microwave use;
however, it can be formed from coaxial cable for use in the UHF and VHF bands.
When made from flexible cable it is easy to bring the desired outlets close together
in order to simplify connections.

18.10.2 The Bagley polygon


This device can be used as a signal combiner or as a power splitter. It can be
thought of as a generalised form of rat race.*
A five-sided form of this circuit is shown in Fig. 18.56. Suppose that a signal is
fed in at O; this will divide into two equal parts going round the loop clockwise and
counter-clockwise. Using arguments analogous to those used with the discussion of
the rat race, it is seen that the signals add at the points A, B, C, D and E giving
equal signals at each. Thus it acts as a five-way power splitter. It will be noted that

* It was devised by G. Bagley of the RAE Antenna Group, Farnborough


924 Coaxial Transmission Lines and Components

alternate points A, C and E are in antiphase to those between them at B and D. In


order to bring all into phase, a length of line X/2 long can be added in the feeds
from B and D.

Fig. 18.57 The Bagley polygon


Characteristic impedance of outputs A — E = Z'o. For system match, the ring to
beofZ0
where Z o = —7=2

where n is the number of matched outputs

If the network is fed by five signals at A, B, C, D and E, each will cancel at each
of the other input points but all will add at 0. Hence provided that the phase of the
sources at B and D have been reversed by an extra X/2 of line, the polygon acts as a
five-way signal combiner.
Similar reasoning can be applied to any polygon with an uneven number of sides.
If it is desired to combine an even number of signals, or to split power into an
uneven number of parts, this can be done by using a polygon of one more side than
the number of parts and then omitting a connection from one of the junctions,
preferably the one opposite to 0 .
The matching of the polygon is achieved when the characteristic impedance of
the quarter-wave lines OA and OE is
Coaxial Transmission Lines and Components 925

Z o = —=• (18.47)

where ZQ = characteristic impedance of the lines connected at A, B, C, D and E and


n = number of loaded junctions.
It is usual to make the X/2 long sides of the same characteristic impedance as the
quarter-wave lines, as this minimises the VSWR on the sides of the polygon.
This circuit can be made of stripline for microwave use. It has the advantage over
other forms of power splitter (or signal combiner) that it does not call for a wide
range of characteristic impedances in order to match. This is because the polygon
ZQ is proportional to \fn for impedance matching. Thus if ZQ is taken as 5012, the
values of Z o for matching, for various values of n, are as shown below:

n = 3 4 5 7 9 11
Z o = 57.7 50.0 44.7 37.8 33.3 30.212

It will be seen that in each of the above cases the Z o of the polygon will not dif-
fer much from 5012 and, when made of stripline, require strip widths differing little
from that used for 5012.

18.11 Plugs and sockets or connectors

These devices are simple in function but can be outstandingly troublesome in


service. This is because all too often unreasonable demands are made upon them:
(a) They are mounted in exposed positions where falling objects land upon them
(b) They are put on light alloy panels and left to withstand the difference in con-
tact potential between the panel and the copper of the cabling
(c) They are expected to withstand violent tugs on their cable mountings and many
other forms of mechanical ill-treatment
It must be remembered that a plug and socket is in fact a piece of precision
engineering, since we require:
(i) It should have a negligible contact resistance between the joined conductors
(ii) It should not allow a significant leakage of radio-frequency signal
(iii) It should introduce at most a small impedance mismatch
(iv) It should not act as a source of intemodulation products
(v) It should be watertight and airtight both within itself and in connection with
any panel or cable on which it is mounted
Bearing these points in mind one must choose a plug and socket with care,
remembering that in this, as in all cases of precision engineering, one has to pay a
fair price for a good product. In particular, it must be noted that just because a plug
926 Coaxial Transmission Lines and Components
A

u
A C

A C

j—WWWW

Fig. 18.58 Otyfer s/?e// connections in plugs and sockets

or socket is sold as being of a given well-designed type, it does not necessarily mean
that it has been made under the costly conditions of controlled manufacture which
ensure products of consistently good quality. Having bought good plugs and
sockets, they must be assembled correctly if the benefits of their quality are to be
enjoyed.
All these considerations become of increased importance at the highest frequen-
cies, and there is much to be said for purchasing those lengths of cable which are
critical in performance from a manufacturer who is prepared to supply as a unit a
length of cable with appropriate plugs and sockets made to a specification such as
MIL-T-81490(AS).
Having obtained good components, the installation should be designed so that
the critical radio-frequency plugs and sockets are mechanically protected and their
environment kept as undemanding as possible.
Considering the requirements of good plugs and sockets we will start with the
electrical contact. If the screening is to be good, it is necessary that there should be
good contact all around the outer conductor; even a perfect contact at one point
on the periphery does not suffice. Thus in Fig. 18.58a is shown a pair of coaxial
lines with outer conductors in contact at A. Continuity tests would show this as a
perfect junction, and yet it would leak signal badly because in effect the two
lengths of coaxial line are joined by a short length of open twin line which radiates
both directly and by way of currents induced on the outside of the coaxial lines.
Coaxial Transmission Lines and Components 927

It is good practice to arrange these contacts as at AA' of Fig. 18.58Z?. Here the
two parts locate at B, the clamping ring forcing both together around the ring of
contact. It is likely that the two parts will be in contact at only three points, but
the narrowness of the gap between B and C will present a low admittance to any
escaping signal.
The inner conductor can give rise to bad contacts, particularly in designs in
which a central pin protrudes. Also where the inner contact element is soldered
onto the inner conductor of the cable there is the danger that this conductor may
become embrittled and crack with subsequent vibration. This is particularly so
when cables having multistrand inner conductors are used.
For satisfactory performance it is imperative that there be good contact between
the cable braid and the housing of the component mounted on it. The principles are
shown in Fig. 18.58c. The housing is counterbored with a locating step at AA and
provided with an internal locking thread. Two metal washers are slipped over the
braid wires which are folded back over the first washer. A locking ring at C holds
the whole assembly, gripping the braid wires between the step in the housing and
the washers, thus ensuring a sound mechanical grip on the braid and a good
electrical contact around the housing.
In some cases a rubber washer is inserted between the washers B and the locking
ring C in order to ensure the sealing of the cable entry. This is bad practice because
there is a chance that, through ageing or thermal cycling, the rubber washer will
yield with the serious consequences that:
(a) The good contact at A is lost with the result that there is signal leakage
(b) The mechanical grip on the braid is relaxed and the cable works loose
(c) As the cable becomes loose mechanical stresses are taken on the inner conductor
and inner contact which may then fail.
The waterproofing of the cable entry is important but it must not be allowed to
put at risk the contact integrity of the cable.
In some systems there is the possibility of interference through the generation of
intermodulation products. These are generated in nonlinear contact resistances, and
the best way of minimising them is to make sure that all contact resistances are very
small. Hence the precautions taken in order to ensure a good screening will help to
control intermodulation.
The generation of intermodulation products in cables and connectors has been
studied by Benson et al4S>46 It has been concluded that the nonlinear effects are
worst in ferrous alloys and those based upon aluminium. The parts in contact
should be held together with sufficient force to break through any oxide film on
the surfaces so that good metal-to-metal contact is established. It has been shown
that in many cases the contacts between dissimilar metals give rise to large inter-
modulation products, particularly when a hard metal is in contact with a softer one,
because in this case the one may flow before the film on the harder contact is
broken. Extreme cleanliness of all interfaces is required.
It is important that the characteristic impedance of the plug and socket when
connected should be close to that of the cable with which it is intended to function.
928 Coaxial Transmission Lines and Components

If it were possible to maintain the diameters of the conductors used in the cables
through the connector and if the same material were used for insulation as in the
cable, all would be well. However, the inner conductor of the plug or socket must
be larger than that of the cable so that it can be bored to accept the cable wire; and
also, in order to maintain a good voltage rating, the mating face is usually made
larger than the cross-section of the cable dielectric. Furthermore, for mechanical
reasons, it is usual to have locating steps on the inside surface of the housing and
upon the inner conductor. In any practical case the necessary changes in diameter
are accomplished by a combination of the arrangements of steps and coned sections
shown in Fig. 18.49.

18.12 References
1 HOWE, G. W. O.: 'The application of transmission formulae to skin-effect problems' /.
IEE, 54,1916, p. 472
2 SCHELKUNOFF, S. A.: 'The electromagnetic theory of coaxial transmission lines and
cylindrical shields', Bell Syst. Tech. I, 13, Oct. 1934, p. 532
3 MOHR, R. J.: 'Coupling between open and shielded wire lines over a ground place,' IEEE
Trans., EMC-9, Sept. 1967, pp. 34-45. Also 'Coupling between lines at high frequen-
cies, IEEE Trans., EMC-9, Dec. 1967, pp. 127-129
4 International Electrotechnical Commission, Geneva, Publ. 96: 'Cables for radiofrequencies'.
96-0: 'Guide to the design of detailed specifications'; 96-1 'General requirements and
measuring methods'; 96-2 'Relevant cable specifications'
5 HOMANN, E.: 'Geschirmte Kabel mit optimalen Geflectschirmen', Nachrichtentech. Zeit.,
21, 1968, pp. 155-161
6 TYNI, M.: 'The transfer impedance of coaxial cables with braided conductors'. EMC Sym-
posium, Wroclaw, Poland, 22-24 Sept. 1976
7 FOWLER, E. P.: 'Some recent work on the screening efficiency of braided and multi-
screened cables'. IEE Conf. Pub. 65, Electrical Interference in Instrumentation, June
1970
8 UK Atomic Energy Standard Specification AESS(TRG)71181: 'Superscreened coaxial
cables for the nuclear power industry; Pt. 1:* 'General requirements and tests'; Pt. 2:
'Cable data sheets*
9 BLACKBAND, W. T.: 'The choice of impedance for coaxial radio frequency cables', Proc.
IEE, 102, Pt, B, Nov. 1955, pp. 804-14
10 DUMMER, G. W. A., and BLACKBAND, W. T.: 'Wires and r.f. cables'. Pitman, London,
1961
11 DAVIS, R., AUSTEN, A. E. W., and JACKSON, W.: 'The voltage characteristics of poly-
thene cables', /. IEE., 94 Pt. Ill, (29), May 1947; and 95 Pt. Ill, (34) Mar. 1948
12 MILDNER, R. C: 'The power rating (thermal) of radio frequency cables', /. IEE, 93 Pt.
Ill, 1946, pp. 296-304
13 BLACKBAND, W. T.: 'Wire-braid screens for r.f. cables', /. IEE, 102, 1955, pp. 363-364
14 SMITH, R. A.: 'Aerials for metre and decimetre wavelengths'. Cambridge Univ. Press,
1949
15 McLEAN, F. C, and BOLT, F. D.: 'The design and use of radio-frequency open wire trans-
mission lines and switchgear for broadcasting systems', /. IEE., 93 Pt. Ill, July 1946,
pp.191-210
16 JASIK, H.: 'Antenna engineering handbook', McGraw-Hill, 1961
17 DUNCAN: 'Characteristic impedance of grounded and ungrounded open wire transmission
lines', Communications, June 1938
Coaxial Transmission Lines and Components 929

18 SHEBES, M. R.: 'Coaxial cable with confocal elliptical cylindrical conductors', Radio-
tecknika, (Moscow) 4, July/Aug. 1949, pp. 36-44
19 OMAR, Y. A. and MILLER, C. F.: 'Characteristic impedance of rectangular coaxial trans-
mission lines' Proc. AIEE, 71, Section T2-75, 1952, pp 81-89
20 SCHELKUNOFF, S. A.: 'Electromagnetic waves', van Nostrand, 1943
21 PARZEN, P.: 'The capacitance/unit length and characteristic impedance of coaxial cables
with one slightly non-circular section', /. Appl. Phys., 18 Aug 1947, pp. 774-779
22 MEYERS, Letter, Elect. Rev. (120), 1942, p. 791
23 ANDERSON, J. C: 'The calculation of characteristic impedance by conformal transfor-
mation', /. Brit. IRE, Jan. 1958, pp. 49-54
24 CHISHOLM, R. M.: 'The characteristic impedance of trough and slab lines', IRE Trans.,
MTT-4, July 1956, pp. 166-172
25 MAGUS and OBERHETTINGER: 'Die Berrechnung des Wellenwiderstandes einer Band-
leitung mit kreisformingem bzw. rechtikigem Aussenleiterquerschnitt', Arch, fur
Elektrotech., 37,1943, pp. 380-390
26 MORTON: 'On the parallel plate condenser and other two dimensional fields specified by
eliptical functions', Phil. Mag., 2, 1926, pp. 827-33
27 SCHELKUNOFF, S. A.: 'Advanced antenna theory', Wiley, 1952
28 THOMPSON, J. J.: 'Recent researches in electricity and magnetism' 1893
29 WILLOUGHBY, E. O.: 'Some applications of field plotting', /. IEE, 93 Pt. Ill, 1946, pp.
275-293
30 WHINNERY, J. R., JAMIESON, H. W. and ROBBINS, T. E.: 'Coaxial line discontinuities',
Proc. IRE, 32, (11), Nov. 1944, p. 695
31 MILES, J. W.: 'Plane discontinuities in coaxial lines', Proc. IRE, 35, Dec. 1947, pp.
1498-1502
32 MEINKE, H. H.: 'Coaxial line elements which produce no disturbing fields at high fre-
quencies', Hochfrequenztech. u. Elektroakust., 61, May 1943, pp. 145-161
33 KRAUS, A.: 'Measured curves of reflection coefficients of compensated inhomogeneities
in coaxial lines and the optimum dimensions deduced from them', Rohde u. Schwarz
Mitt., 8, Dec. 1956, pp. 1-12
34 WILLIS JACKSON: 'High frequency transmission lines', Methuen, London, 1945
35 WHEELER, H. A.: 'Transmission lines with exponential taper', Proc. IRE, 27, 1939,
pp.65-71
36 CHRISTIANSEN, W. N.: 'An exponential transmission line employing straight conductors',
AWA Tech. Rev., 7, 1947, pp. 229-40
37 NIUTTA, A.: 'Impedance matching using a special type of exponential line', Poste e Tele-
comm., 17, Aug. 1949, pp. 417-23
38 ROHRMAN: 'Improvement of the transformation properties of the exponential line by
compensation arrangements', Arch. Elektr. Ubertr., 4, Jan. 1950, pp. 23-31
39 GURLEY, J. G.: 'Impedance matching by non-uniform transmission lines', IRE Trans.,
PGAP-4, Dec. 1952, pp. 107-109
40 KLOPFENSTEIN, R. W.: 'A transmission line taper of improved design', Proc. IRE, 44,
Jan. 1956, pp. 31-35
41 GENT, A. W. and WALLIS, P. J.: 'Impedance matching by tapered transmission lines', /.
IEE, 93 Pt. IIIA, 1946, p. 559-563
42 MARCHAND, N.: Transmission line conversion', Electronics, 17, 1944, pp. 142-145
43 CLOETE, J. H.: 'Exact design of the Marchand balun', Microwave J., May 1980, pp. 99
and 102
44 CLOETE, J. H.: 'Graphs of circuit elements for the Marchand balun', Microwave /., May
1981,pp. 125-128
45 BAYRAK, M. and BENSON, F. A.; 'Intermodulation products from non-linearities in
transmission lines and connectors at microwave frequencies', Proc. IEE, 122, 1975,
pp.361-367
46 AMIN, M. B. and BENSON, F. A.: 'Non-linear effects in coaxial cables at microwave
frequencies', Electron. Lett., 13, 1977, pp. 768-770
Index

AWACS low sidelobe arrays, 163 Sidelobe cancellers, 400-407


Acoustic arrays, 55 Single loop cancellers, 400-404
Active antenna, 639-640,686,697 Adcock array, 831-834
Active impedance, 177-179 Aerodiscone, 738
Adaptive arrays, 331, 407-454 Airborne radomes, 507-536
Analogue feedback arrays, 442- Abberation, 515-520
444 Cross-polarisation, 521-528
Binomial amplitude weighting, Dielectric rings, 518
414 Lightning strike protection, 536
Broadband jamming, 449-453 Multi-band designs, 511-515
Cascade processors, 449 Pilot-static system, 534-536
Digital implementation, 444-449 Rain-erosion cap, 533-534
Eigenbeam analysis, 435-441 Reflections from radome, 528-
Eigenbeams, 421-427 532
Eigenvalues for jammers, 427- Wide-band transmission designs,
435 507-511
Fifteen element array, 444 Wire compensation, 518
Four element array, 414, 438, Aircraft antennas, 625,637, 713-
439,452 716,794-819
Gradient descent method, 445 Air-traffic control, 802-805
Jammers, 427-435 Caps, 715
Least mean square algorithm, 446 Clothes-line, 713
Main beam constraints, 417-419 Communications, 796-801
Matched filter criterion, 435 Coverage requirements, 794-796
Matrix inversion, 412-417 Dipole-director pair, 816
Mobile communication antenna, Direction Finding ratio, 813
416,417 Glidepath antennas, 809-810
Mutual coupling effects, 411, 422 HF antennas, 713-716
Retrodirective beams, 419-421 Homing antennas, 812-817
Sample matrix inversion tech- Instrument Landing Systems,
nique, 447 805-806
Seventeen element array, 444 Loop (terminated), 816
Signal-to-noise criterion, 408-412 Marker antennas, 810-811
Adaptive null steering, 399-454 Monopoles, 797-805, 807, 814
Adaptive arrays, 407-454 Navigation, 805
Bandwidth, 404 Noise at MF, 637
Cancellation at IF, 407 Non-metallic structures, 818
Multiple loop, 404-407 Nose dipole, 808
Index 931
Notch, 714, 804 Phasers, 166
Notch fed plate, 807 Power dividers, 98-100
Parasitic elements, 762 Radial line, 166
Propellor or rotor modulation, Resonant arrays, 81-86
819 Resonant grid, 164
Ramshorn, 807 Resonant-slot design, 84-86
Slots, 715, 803, 809, 816 Ring hybrids, 164
Telemetry and command, 805 Scanned array, 103
Trailing wire, 715 Series feed, 79, 166
Transmission line antenna, 800, Series-parallel feed, 167
810 Series-series feed, 167
VLF/LF wire antennas, 625 Serpentine feed, 96
VOR antennas, 811 Serrated guide, 81
Whips, 715, 800 Shunt, 79
Wing-tip antennas, 808 Single-beam condition, 87
Wire-grid modelling, 713 Slab line feed, 162
Yagi, 816 Slot array design, 93-96
Yagi array, 810 Snake feed, 96-98, 166
Alford loop, 747,847 Squint feed, 96
Amplitude-comparison monopulse Stripline, 100
antenna, 333 Stripline slot, 165
Anti-fading antennas, 608-611 Sub-array lobes, 104-106
Aperture array distribution, 2 TEM line coupler, 100
Aperture excitation efficiency, 157 Travelline wave arrays, 87-98,
Apertures (cylindrical array of), 240- 161,165
257 Waveguide (periodically loaded),
Appleton layer, 665 98
Areal beamwidth, 147 Waveguide corporate feed, 161
Array excitation, 141-161 Wire grid array, 164
Array feeds Array lattice, 141-143
Array filling time, 103-104 Array signal processing, 330-456
Azimuth power divider, 164 Adaptive arrays, 407, 454 (see
Bifurcated T waveguide, 98 Adaptive arrays)
Branch-line phaser, 101 Adaptive null steering, 399-454
Coaxial-line T junctions, 98 (see Adaptive null steering)
Corporate feed, 79, 98-101, 166 Applications, 331-332
Cross-guide, 161 Array thinning, 344-346
Digital phasers, 106-108 Code separation of elements, 352
Distributed array, 79, 102 Code separation of transmitters,
Electronic scanning, 166-169 351
Element conductance, 90-92 Digital and baseband beam form-
Fixed beam, 161-166 ing, 356
Frequency scanning, 96-98 Directionally decorrelated beams,
Frequency squint, 87 349-351
Helix, 98 Doppler (pulse), 342
Hybrid junctions, 98 Doppler VOR beacons, 337
Hybrid monopulse network, 161 Frequency filtering, 340
Image element array Frequency scanning at AF, 341
Linefeeds, 101-102 Frequency separation of elements,
Microstrip elements, 81, 100 352
Parallel-parallel feed, 168 Frequency separation of trans-
Parallel-series feed, 167 mitters, 351
Periodically loaded line, 81 Interferometers, 357-360
932 Index
Intermediate frequency process- Baluns, 727-728, 917-922
ing, 355 Bats wing antenna, 752
Low sidelobe tapers, 340 Beam cophasal excitation, 299-304
MTI, 342 Beam efficiency, 35-36
Mills cross array, 345, 357 Beam forming network (orthonor-
Multiple beams, 332-341, 347 mal), 391-395,434
Multiplicative arrays, 360-371 Beverage antenna, 682-683
Multiplicative signal processing, Bickmore-Spellmire array distri-
357-371 bution, 37-39
Multistage systems, 353 Bicone antenna, 728
Null steering, 345, 385-454 (see Blass matrix, 334
Null steering) Blind spot in arrays, 177
Null steering arrays, 388-395 Bow-tie antenna, 686
Nyquist jate sampling, 337-341 Broadcast antennas, 606-624
Pulse compression in angle, 343- Anti-fading antennas, 608-611
344 CBC station CHFA, 612
Radio astronomy (use in), 357 CFTR station, 614
Receiving systems for radar, 332- Directional arrays, 611-615
342 Folded monopole, 607
Sidelobe canceller, 331, 395- Franklin antenna, 610
397 Grounded tower antenna, 609
Single sideband modulation, 338 High-rise building influence, 620
Squirt technique, 332 J-antenna, 609
Storage filter, 338 LF antennas, 607-608
Synthetic (coherent) aperture MF broadcasting stations, 608
antennas, 371-385 Power line influence, 617, 620-
Synthetic aperture techniques, 624
357-371 (see synthetic Re-radiation distortion of MF
aperture antennas) antennas, 615-624
Technologies, 354-356 Bruce antenna, 679
Thinned array, 364 Butler matrix, 314-316, 319, 324,
Time separation of elements, 352 334,355,356,398,407,
Transmit and receive antenna 421
systems, 342-356
Within pulse frequency scanning,
348-349 Canonical minimum scattering
Within pulse scanning, 343-344 antenna, 179-180
Within pulse time sampling, 347 Charge coupled devices, 355
Array thinning, 344-346 Circular arrays, 142, 156, 298-329,
Arrays, see Adaptive arrays, Array 834
feeds, Array signal pro- NRL 32-element array, 322
cessing, Conformal arrays, Amplitude modes, 306-307
Conical arrays, Cylindrical Applications, 299, 319-327
arrays, HF arrays, Linear Bandwidth, 308, 313
arrays, Mutual coupling Beam cophasal excitation, 299-
in arrays, Planar arrays, 304
Superdirective arrays, Beamwidth, 308-309
Thinned arrays, Yagi-Uda Butler matrix, 314-319, 322, 324
arrays Communication applications,
319-324
Commutated direction finder,
Babinet flat strip dipole, 181 323-324
Backfire (short) antennas, 758-759 Concentric, 698-699
Index 933

Dipoles,312,321,323 Junctions, 914-917


Direction finding applications, Leakage, 862
298,299,323-327 Mutual screening, 912-914
Direction elements, 310-314 Non-circular inner conductor,
Disc array, 142 882
Electronic beam rotation, 300 Operating principles, 860
Electronic scanning, 302-304 Power dividers, 922-925
Excitation networks for phase Screened twin lines, 868-871
modes, 314-318 Screening, 862
Goniometer, 697,699 Semi-circular outer, 880-881
HF arrays, 302, 319, 324,697- Square outer, 880-881
699 Stranding effects, 881-882
Inter-element spacing, 309, 312 Thermal ratings, 868
Lens feeding, 323 Voltage ratings, 868
Logarithmic, 708 Cone antenna, 738
Low sidelobes, 300, 313 Conformal array analysis, 229-240
Luneberg lens, 302, 320 Asymptotic methods, 230
Maxson matrix, 317 Beam ports, 231
Microwave arrays, 324-327 Characterisation of an array, 230-
Monopoles, 316, 319, 323,697 233
Mutual coupling, 298, 318-319 Feed network model, 231
Null steering, 398 Free excitation, 232
Null steering array, 320 Gain (realised pattern), 232
Omnidirectional elements, 309- Modal methods, 230
310 Multimode aperture elements,
Pattern synthesis, 306-307 235-237
Phase modes, 305-306, 309-318 Mutual admittances, 238-240
Position finding, 834 Mutual coupling (Internal and
Radar applications, 321-323 External), 233-235
Radius (choice of), 307-308 Mutual coupling, 232
Signal processing techniques, Radiation network model, 231
303-304 Reflection coefficient (Active),
Slot radiators, 312 234
Superdirectivity ,321 Scattering coefficients, 232, 235-
Wullenweber arrays, 299, 301, 240
319,324 Waveguide elements, 235
Clothes-line antenna, 713 Conformal arrays, 227-297
Coaxial transmission lines, 860-929 Analysis, 229-240, (see Con-
Attenuation, 861, 863-867 formal array analysis)
Baluns, 917-922 Aperture on a conducting
Connectors, 925-928 cylinder, 290-295
Coupling between, 906-912 Applications, 227
Coupling with a fusalage, 909- Beam ports, 231
912 Conical arrays, 274-284, (see
Elliptical conductors, 880 Conical arrays)
Frequency limitations, 867 Cylindrical array of apertures,
Impedance (characteristic), 863- 240-257
865,870 Cylindrical arrays, 257-270, (see
Impedance (transfer), 861, 906 Cylindrical arrays)
Impedance-matching sections, Definition, 227
916-917 Feed system design, 229
Interference, 912-914 Fock functions, 246, 255, 257,
Intermodulation, 927 288-290
934 Index
Fourier transforms for wave- Eigenpatterns, 258-261
guides, 286-288 Eigenvalues of admittance
GTD evaluation of mutual admit- matrices, 261-264
tance, 270-274 Element pattern design, 265-267
Generic convex surface, 270-274 GTD expressions for mutual
Hybrid junctions, 235 admittances, 247-252
Modal admittance for waveguides, GTD surface field, 244-247
284-285 Lit far-field, 255
Multimode aperture elements, Magnetic current on a cylinder,
235-237 240-242
Mutual admittance, 270-274 Numerical examples, 267-270
Periodic cylindrical array, 295- Penumbra far-field, 255
296 Periodic array model, 257-258
Periodic structure analysis, 257- Radiation pattern evaluation,
270 252-257
Polarisation effects, 228 Realized gain pattern, 264
Power dividers, 235 Reciprocal lattice, 263
Reduction of boresight error, 228 Scan coverage, 268
Conical arrays, 274-284 Shadow far-field, 255
Admittances, 274-278 Slots (circumferential), 247
Cone tip scattering, 277 Watson transformation, 253
GTD formulation, 278
Radiation pattern, 278-284 Diamond antenna, 586
Conical transmission lines, 887-901 Dielectric covers for antennas, 842-
Angular strip on plane, 897 844
Cone on edge of a sheet, 898 Dielectric-loaded antennas, 840-
Conical sheet normal to plane, 844
894 Ferrite loading, 840-841
Cylinder parallel to thin sheet, Log-periodic, 841
898 Monopoles, 841
Double conical sheet, 894 Rectangular cavity slot, 841
Impedance calculations, 904 Yagi, 841
Pair of equal parallel strips, 897 Dipoles
Split-cylinder, 895 Aircraft (use on), 797-805, 807,
Stereographic projection, 888 814
Strip between two planes, 896 Arrays of, 16-20, 85, 109-113,
Strip parallel to plane, 896 186-193,689-697,764-
Two conical lines, 891 767,774
Corner antennas, 755-758 Axially fed, 731
Corporate feed, 79, 98-101,166 Babinet flat-strip, 181
Critical frequency, 664 Baluns, 919
Cubical quad antenna, 766 Bats wings, 752
Current element, 673 Bicone, 728
Current sheet, 172-175 Bow-tie, 686
Cylindrical arrays, 240-270, 298- Broadband, 728
329 Cage, 686
Admittances (mutual), 242-252 Circular arrays, 312, 322, 323
Admittances (self), 242-244 Coaxial, 732-733, 792, 793
Aperture matching, 265-267 Coaxial bicone, 733
Apertures, 240-257 Collinear array, 16-20
Arbitrary excitation, 264-265 Coupling between, 181-221,
Eigenexcitations, 258-261, 295- 850-853
296 Crossed, 111,113,748,847
Index 935
Curtain of, 691-693 Director element, 763
Cylindrical, 183 Discone antenna, 733, 738
Directivity of linear array, 16 Diversity reception, 669
Director pair, 816 Dolph-Chebyshev distribution, 20-
Disc, 761 25,63-67
Discone, 733,738 Dolph-Chebyshev sidelobes, 156
Equivalent to slots, 85 Doppler VOR beacon, 332, 337, 838
Fat, 728 Doublet antennas (coupling
Feeds for reflectors, 754-762 between), 852
Folded, 703,728 Doublet elements in HF arrays, 695
Full wave and tuned reflector,
764 Effective aperture, 12-20
Gain standard at HF, 672 Electronic Support Measures (ESM),
HF arrays, 689-697 298,327
Inverted-V,634 Electronic counter measures, 385
Loaded, 704 Electronic scanning array, 4,166-
Logarithmic arrays, 700-706 169
Masts (effects of), 753 Extraordinary wave, 664
Mechanical design at HF, 702-
Feeds
721
Array (see Array feeds)
Mobile applications, 625-627
Dipole-disc, 761
Mutual coupling between, 181-
Dipole-reflector, 760-762
221,850-852
Slotted dipole, 761-762
Mutual resistance between, 17,
Ferrite loaded antenna, 840-841
154
Ferrite loop antenna, 637
Nose, 808
Fish-spline reflectors, 757
Parallel array, 16-20
Fishbone antenna, 690
Parallel to cylinder, 786
Flat plane array, 154, 161-166
Parasitic (use as), 762-767
Fock functions, 288-290
Parasitic as reflector, 763-767
Franklin antenna, 556,603,610
Position finding (use for), 837
Printed circuit, 113 Geometric Theory of Diffraction
Printed slot-dipole, 850 244-252,270-274,278,
Reflector, 760-762 536,550,779-780
Resonant arrays, 85 Goniometer, 697,699
Scale-modelling at HF, 718 Grating lobes, 9-12,147-151
Short (current line), 122 Hexagonal lattice, 149-151
Sleeve, 727 Rectangular lattice, 147-149
Slotted, 761-762 Ground radomes, 536-549
Stacked elements, 749-753 Ground wave, 553
Super-turnstile, 752 Ground wave propagation, 668
TV transmitting, 752-753 Grounded antenna, 554
Turnstile, 748, 751
Vee type, 769 HF antennas, 663-724
Vertically polarised, 727-734 Active, 686,697
Direction finding (DF) systems, 298, Aircraft antennas, 713-716
299,314 Arrays, 689-699, (see HF arrays)
Direction-finding antennas, 688, Beverage antenna, 682-683
831-837 Bruce antenna, 679
Directivity Coaxial feeders, 722
Arrays of dipoles, 16-20 Design techniques, 717
Linear arrays, 12-20 Dipole antennas, 686-687
Planar arrays, 151-156 Horn, 708-709
936 Index

Log-periodic, 699-708 E layer reflection, 665


Logarithmic arrays, 706-708 Extraordinary wave, 664
Logarithmic dipole, 700-706 F layer reflection, 665
Long-wire antenna, 677-678 Fast fading, 669
Loop antennas, 687-689 Focussing, 670
Masts., 721 Frequency diversity, 669
Measurements, 719 Ground characteristics, 672-677
Mechanical design, 720-722 Ground dielectric constant, 672
Mobile applications, 711-717 Ground fresnel zone, 675
Monopoles, 683-686, 711 Ground improvements, 675-677
Non-resonant antennas, 677-683 Ground reflection coefficients,
Propagation, (see HF Propaga- 673-674
tion) Ground reflections, 612-615
RCA Fishbone, 690 Ground screens, 675-677
Receiving, 301 Ground wave propagation, 668
Rhombic antenna, 680-682 Ionospheric conditions, 663-668
Scale-modelling, 717-718 Lowest usable frequency, 668
Siting considerations, 722-723 M reflections, 666
Sloping-vee antenna, 679 Maximum usable frequency, 666
Slot antennas, 686-687 Multipath distortion, 669
Transmission lines, 722 Multipath propagation, 666
Travelling wave antenna, 677-678 Ordinary wave, 664
Vee antennas, 679-680 Polarisation diversity, 669
Whips, 686, 711 Polarisation of wave, 664
Wire grid lens, 709-711 Random differential phase error,
Wire grid modelling, 717 669
HF arrays, 689-708 Signal strength, 671
Broadside, 691-697 Signal to noise ratio, 670
Circular, 319,324,697-699 Single hop wave, 666
Curtain of dipoles, 691-693 Skip distance, 666
Dipoles, 689-691 Sky wave propagation, 663
Doublet elements, 695 Slow fading, 669
Endfire, 689-691 Solar activity, 665
Inverted-Vee, 6 97 Space diversity, 669
Koomans, 691 Thermal noise, 670
Logarithmic, 699-708 Time diversity, 669
Monopoles, 694,697 Half-delta loop antenna, 636
Phased, 693 Hamming array distribution, 39
RCA Fishbone, 690 Hansen one-parameter distribution,
Rhombics,697 159-161
Series phase, 690 Hansen-Woodyard array distribution,
Wullenweber,697 56-61
HF propagation, 663-672 Helical antennas, 770-772, 837
Angle diversity, 669 Helical monopole, 739
Angle of fire, 673 Helically loaded antenna, 629
Antenna design, 666 Helix (use as array feed), 98
Antenna directivity, 670 Helix inductors, 577
Antenna impedance match, 672 Horns
Attenuation loss, 664 Coupling between, 185-186
Brewster angle, 675 HF tapered aperture, 708-709
D layer absorption, 668 Hula loop, 739
D layer reflection, 665 Huygens source, 175
Diversity reception, 669 Hybrid junctions, 98
Index 937

Conformal arrays, 235 Element pattern, 2, 16


Micro strip, 100 Elements, 109-133
Ring, 100 Endfire, 309
Short-slot, 98 Errors in, 73
Signal processing, 330 Excitation (forced) (definition), 2
Strip line, 100 Excitation (free) (definition), 2
Top-wall, 100 Excitation design, 2-79
Excitation efficiency, 29
Ideal element pattern, 175 Feeding network, 15
Image element array, 165 Feeds, 79-108, (see Array feeds)
Impedance crater, 194 G/T, 50
Interferometers, 357-360, 835-837 Gain limits, 73-79
Interferometric arrays, 331 Grating lobes, 9-12
Inverted-Vee array, 697 Hamming distribution, 39
Ionosphere, 663-668 Hansen-Woodyard distribution,
Appleton layer, 665 56-61
D layer, 665 Integrated elements, 132
E layer, 665 Low sidelobe designs, 39-41
F layer, 665 Low sidelobe patterns, 4-42
Kennelly-Heaviside layer, 665 Micro strip patches, 127-130
Moderate gain elements, 133
J-antenna, 609-610 Monopoles, 110, 113
Josephson's monopole, 736 Multimode waveguide elements,
123-126
Koomans array, 691 Mutual coupling, 54, 170-221,
(see Mutual coupling in
LF antennas (see VLF, LF and MF arrays)
antennas) Narrow beam patterns, 4-42
Leaky wave antenna, 81 Pattern synthesis, 42-54
Leaky wave in array, 215 Patterns, 5-6
Linear arrays, 1-140, 170-221 Printed circuit elements, 126-133
Adaptive arrays, 407-454, (see Probability density function, 74
Adaptive arrays) Quantisation lobes, 104-108
Aperture distribution, 2 Scan compensated elements, 121—
Aperture efficiency, 31 126
Array elements, 109-133 Shaped beams, 42-45
Array-factor pattern, 16 Sidelobe adjustment, 36-37
Baffles, 123 Sidelobe envelope shaping, 45-48
Beam efficiency, 35-36 Sidelobe level, 4, 41-42, 74
Beamwidth,6-8, 25,33 Sidelobe ratio, 4
Bickmore-Spellmire two- Sidelobes, 8-9
parameter, 37-39 Signal-to-Noise Ratio, 48-53
Circular arrays, 322 Slots, 16-20, 113-121
Constrained synthesis, 48-54 Space factor, 2
Dipoles, 16-20,109-113,122 Spatial filter, 124
Directional elements, 12 Stripline slots, 126
Directivity, 12-20, 25, 29 Sub-array lobes, 104-106
Discrete Fourier Transform, 2 Superdirectivity, 23, 48, 54-68,
Dolph-Chebyshev distribution, (see Superdirective arrays)
20-25,63-67,76-80 Tapered excitation, 20-42
Edge effects, 190 Taylor distribution, 25-37, 76
Effective aperture, 12-20 Thinned, 68-73, (see Thinned
Element impedance, 190-193 arrays)
938 Index
Tolerances, 73-79 Modal methods, 230
Travelling wave patch arrays, Monopoles
130-133 Active, 686, 697
Uniformity excited, 3, 5-20 Aero disco ne, 738
Unit cell concept, 123 Aircraft (use on), 797-805, 807,
Unit circle, 3 814
Woodward and Lawson syn- Annular, 737,743-745
thesis, 43 Aperiodic, 754-762
Lit far-field, 255 Array elements (use as), 110, 113
Log-periodic antennas, 676, 699- Bow-tie, 113
708,717,718,720,767- Broadband, 735
769,848 Circular array, 316, 319, 323
Arrays, 706-708,717 Circularly polarised, 847, 849
Broad-dipole, 703 Cone, 738
Circular, 708 Current distribution, 603
Design, 704-706 Dielectric loaded, 841
Dielectric loaded, 841 Discone, 738
Dipoles, 700-706 Double cone, 684
Dual polarised, 701 Dual-band, 854
HF, 699-708, 720 Electrical height, 564
History, 699 Elevated feed, 683, 694, 697
Interleaved array, 707 Folded, 607
Loaded dipole, 704 Grounded, 554, 569-574
Planar, 767-769 HF, 683-686, 711
Pyramidal, 769 Helical, 739
Scale modelling at HF, 718 Hula loop, 739
Toothed, 769 Inverted-L,684,849
Trapezoidal, 769 Inverted-T, 684
Zig-zag, 769 Josephson, 736
Long-wire antenna, 611-618 Low profile, 737,739-743
Loop antennas, 637,687-689, 747, Mechanical design at HF, 721
751,762,816 Mutual impedance between,
Lowest Usable Frequency, 668 565-569
Luneberg lens, 302, 320, 709 Near conductor, 783
Notched plate, 739
MF antennas (see VLF, LF and MF Pill box, 737
antennas) Planar, 737
MF ground loss resistance, 573 Printed, 736
Man as an antenna, 824-827 Quarter-wave, 734
Marine vehicle antennas, 716-717, Reduced height, 735
793-794 Series inductance loading, 683-
Maximum Usable Frequency, 666 684
Maxson matrix, 355 Short, 790
Method of Moments, 559, 778 Sleeve, 734-735, 789, 855
Micro strip patch arrays, 81, 100, Top-loaded, 560, 683-684, 735
127-130,165 739,792
Mutual impedance between, Transmission4ine, 737, 739-743
184-185 Umbrella, 684
Mills cross array, 345, 357, 836 Vehicles (use on), 783, 789, 790,
Minimum redundancy array, 365 792,819
Minimum scattering antenna, 179- Vertical, 554-577, 598-603
180 Motor vehicle - parasitic elements,
Mobile antennas (see Vehicular 762
antennas) Moving Target Indicator (MTI), 342
Index 939
Multiband VHF & UHF antennas, Notches on aircraft, 714
854-856 Null steering, 345, 385-454
Multiple beam antennas, 331,332- Adaptive, 407-454
341,347-352 Beam-forming network, 391-393
Multiple-tuned VLF antennas, 597- Multi-element arrays, 388-395
598 Multiple nulls, 386-388
Multiplicative arrays, 360-371 Mutual coupling effects, 398
Mutual coupling in arrays, 170-221 Open4oop, 393
Active element efficiency, 197- Sidelobe canceller, 395-397
200 Null steering antennas, 331
Active element pattern, 177-179, Null steering arrays, 320, 388-395
189,218-221
Active impedance, 177-179 Open-wire transmission lines
Babinet flat-strip dipoles, 181 Balanced two-wire above earth,
Blind spot in arrays, 177 878
Blindness, 212-217 Conductors in enclosures, 883
Canonical minimum scattering Corona rings, 872
antenna, 179 Eccentric circular, 876-877
Carter dipole formula, 85, 183, Five-wire, 879
184 Four-wire, 878, 879
Current sheet, 172-175 Impedance, 871-880
Dipoles, 181-221 Impedance calculations, 902-904
Element-by-element approaches, Peek's law, 872
181-203 Power-handling, 872
Forced excitation model, 175- Six-wire, 879
176 Striplines, 883-887, 892-901
Free excitation model, 176-177 Three-wire, 878
Grating lobe series, 193-197 Two-wire, 871-876
Grating lobes, 201-203 Two-wire in parallel above
Horns, 185-186 ground, 879
Ideal element pattern, 175, 197 Wire in corner, 878
Impedance crater, 194 Ordinary wave, 664
Impedance matching networks,
200-203
Impedance matric solution, 186-
193 Parasitic elements, 762-767
Microstrip patch elements, 184- Pawsey stub, 728
185 Penumbra far-field, 255
Minimum scattering antennas, Phased arrays, 4, 172, 230, 693
179 Pill box antenna, 737
Periodic cell approach, 203-217 Planar arrays, 141, 226
Scan compensation, 218-221 Aperture excitation efficiency,
Simulators, 217 157
Slots, 193-198 Array aspect ratio, 144
Wire antennas, 181 Beam scanning, 147
Beam width, 143-147
Brick waveguide array, 215
Circular disc, 142
Notch antennas, 714, 739, 745-747, Corrugations, 218
762,804,807,855 Cross plane, 144
Dielectric cover for, 844 Dipoles, 181-221
Disc, 747 Directivity, 151-156
Multiband, 855 Distributed waveguide array, 169
Parasitic (use as), 762 Dolph-Chebyshev sidelobes, 156
940 Index

EAR array, 169 Bagley polygon, 923-925


Effective aperture, 155 Rat race, 922-923
Electronic scanning, 166-169 Printed circuit arrays, 355
Element pattern, 161 Printed circuit elements, 126-133
Excitation, 141-161
Feeds, 161-169, (see Array feeds) Quantisation lobes, 104-108,151
Ferrite scanned edge slot, 166 Quartz, 478, 508, 511,672
Fixed beam arrays, 161-166
Flat plane, 154,161-166 RASSR array, 169
Frequency scanned edge slot, 166 RCA Fishbone array, 690
Grating lobes, 147-151 Radars
H-plane baffles, 218 Airborne mapping, 371, 382
Hansen one-parameter distri- Doppler (pulse), 418
bution, 159-161 H.F.,371
Lattice, 141-143 Monopulse,332
Minimum scan angle, 227 Pulse-compression, 418
Mutual coupling, 170-221 Satellite mapping, 371, 382
Printed circuit spiral array, 169 VHF synthetic aperture, 385
Quantisation lobes, 151 Radiation resistance, 559
RASSR array, 169 Radio astronomy, 331, 357
Reciprocal cell, 260 Radio refractive index, 663
Scan compensation, 218-221 Radio telescopes
Scan plane, 144 Brazilian, 4 71m
Side-lobe ratio, 144 Five Colleges Astronomy Obser-
Simulators, 217 vatory, 476m
Slot with parasitic element, 218 Swedish, 476m
Slots, 181-221 Radome induced boresight error, 228
Space factor, 142, 143 Radome materials
Stripline slot, 165 Alkyd triallycyaniviate copoly-
Taylor distribution, 156-159 meric, 487
Transition angle, 144 Alumina, 463, 478,488, 489-
Waveguide linear, 154 492,499-500
Waveguide slots, 142 Anisotropic dielectrics, 519
Wire grid, 164 Beryl oxide, 490
Plasma frequency, 663 Beryllium, 495
Position-finding antennas, 830-839 Boron nitride, 490, 495
Adcock array, 831-834 Cordierite, 490, 495
Circular arrays, 834 Core materials for resin fibres,
Commutated antenna system, 834 483-488
Commutational VOR, 838 Dacron,476,541
Dipoles, 837 Dielectrics (artificial), 498-499
Direction finding systems, 831— Epoxy laminates (properties),
837 479-483
Doppler VOR, 838 Esscolam VI, 546-548
Helices, 837 Fibre reinforced resin laminates,
Homing systems, 830-831 478-483
Interferometer, 835-837 Foams, 485-488
Mills Cross, 836 Glass aluminium phosphate, 495
Navigation systems, 837-839 Glass ceramic, 478, 488, 492-494
Orthogonal arrays, 834 Glass fibre laminate, 463-466,
Satellite tracking, 836-837 476,500,508
Tracking systems, 831-837 High temperature, 488-498
Power dividers, 922-925 Honeycombs, 483-485
Index 941
Isocyanate, 486 Ground radomes, 536-549
Metallic elements (inclusion of), History, 457
510-511 Ice protection, 539-541
Mullite, 495 Inflated, 542
Neoprene coating, 483,498-502 Lightning protection, 539-541
Nitrile ebonite, 485 Manufacturing methods, 479
Nomex,483 Materials (see Radome Materials),
Nylon, 483 476-501
PTFE, 476 Mismatch effects, 528
Phenolic, 483 Missile, 478
Polyester laminate, 478, 479- Moisture effects, 537-539
483,508 Multiband performance, 467
Poly imide resins, 483 Near-spherical rigid shapes, 501-
Polyols, 486 503
Polysethersulphone, 476 Noise temperature effect, 539
Polyurathane coating, 483, 500- Non-spherical rigid shapes, 503-
502 507
Polyurathane foam, 485-487 Rain effects, 537-539
Pyroceram, 490, 499-500 Rain erosion protection, 483
Quartz, 478, 508, 511 Ray tracing, 518
Rain-erosion resistant, 499-501 Reflections from, 528-532
Requirements, 476-478 Requirements for radomes, 461-
Rubberised fabric, 476,541 462
Silica, 489,490,494-495 Rigid, 542-549
Silicon nitride, 489,490,495, Shaped radome design, 469-475
500 Ship borne radomes, 536-549
Silicone-phenolic resin, 483 Single layer, 458,463, 507-532
Spinel, 495 Slot waveguide array, 161
Streatite,495 Snow protection, 539-541
Syntactic foam, 488 Space frame, 503, 537,541, 543-
TDI, 487 549
Tedlar, 546-547 Structures, 501-506
Radomes, 457-552 Transmission properties, 462-469
A-type sandwich, 458, 464-466, Voltage breakdown ,528
507-532 Wide pressure effects, 537
Advantages, 460-461 Ramshorn antenna, 807
Air supported, 541-542 Rat-race, 922-923
Airborne applications, 503-507 Ray tracing techniques, 779-780
Airborne radomes, 507-536 Realized gain pattern of an array,
B-type sandwich, 459,466-467, 232, 264-265
507-532 Reflector element, 763
Blind spot in array, 215 Reflectors
C-type sandwich, 459,468-469, Aperiodic, 754-762
507-532 Backfire, 758
Computer-aided design, 473, 550 Corner, 755-758
Constructions, 458-459 Cylindrical parabolic, 98
Curved, 525-527 Dipole feeds, 760-762
Disadvantages, 460 Fish-spline, 757
Electrical properties, 462-469 Flat sheet, 754-755
Environmental requirements, Parabolic cylinder, 760
461-462 Paraboloid, 760
Equivalent transmission line, 462 Parasitic element, 763
GTD(useof), 550 Radomes, 536
942 Index

Shaped, 759-762 Stripline, 126


Vertex mismatch, 761 Vehicles (use on), 820
Remote sensing, 385 Waveguide, 113-121,183
Rhombic antenna, 133,680-682, Slots in stripline, 183
697,706,708,717,720 So dp ole antenna, 850
Rotmanlens, 355 Sonar antennas, 332
Space factor, 2,142,143
Space-frame radome, 503, 537, 541,
Satellite communication antennas, 543,549
796,801 Standing wave feed, 80
Satellite tracking (VHF) antennas, Stick array, 161
836-837 Stripline slot arrays, 165
Scale-modelling, 713, 717-718, 781 Striplines, 883-887, 892-901
Scanned arrays, 204 Superdirective arrays, 23,48, 54-68,
Shadow far-field, 255 201,321
Ship antennas (HF), 716-717 Array superdirectivity, 56-63
Ship borne radomes, 536-549 Bandwidth, 63-64
Skip distance, 666 Directivity, 56-62, 66-68
Skywave,553,663 Dolph-Chebyshev arrays, 63-67
Sleeve antenna (dual band), 855 Efficiency, 63-64
Sloping-Vee antenna, 679 Element superdirectivity, 55-56
Slots Hansen-Woodyard distribution,
Admittance, 115-121 56-61
Aircraft (use on), 715, 803, 809, History, 54
816 Tolerances, 63-64
Annular, 737, 743-745 Surface acoustic wave filters, 355
Broadwall, 113 Synthesis
Cavity (with), 841,848 Butterworth function, 43
Circular arrays, 312 Chebyshev polynomials, 43
Circumferential, 247 Constrained, 48-54
Crossed, 121,848 Iterative sampling, 45
Cylinder, 749, 751, 772-773 Least-squares, 45
Dielectric cover for, 843 Linear arrays, 42-54
Dipole (parallel), 849 Minimax, 45
Dumbbell, 849 Phase-only optimization, 53-54
Edge, 113 Shaped beams, 42-45
H-slot, 121 Sidelobe envelope shaping, 45-48
HF, 686-687 Woodward and Lawson, 43
Iris excited, 121 Synthetic aperture antennas, 331,
Linear arrays, 16-20, 85, 113- 357-371
121 Applications, 385
Monopole (with parasitic), 849 Coherent, 371-385
Mutual coupling between, 181- Doppler frequency interpretation,
221 382-383
Mutual resistance between, 154 Focused, 376-380
Parasitic on aircraft, 765 Motion (response to), 384-385
Pilon, 773 Principles, 371-374
Planar arrays, 142, 161,181-221 Radar PRF, 382
Printed slot-dipole, 850 Radar performance, 383-384
Probe coupled, 121 Samples (minimum number),
Resonant arrays, 85 380-382
Series, 113 Signal processing hardware, 382
Shunt, 113 Target scintillation, 384-385
Index 943
Unfocused, 374-376 Dipoles, 727-734, 748-767, 769,
774,786,792,793,797-
T-type antenna, 580-584 805,807,808,814,816,
TV transmitting antennas, 752-753, 837,847,850-853
773-775 Directional antennas, 753-775
Taylor circular n distribution, 156- Frequency range, 725
159 Helical, 739, 770-772, 837
Taylor source distributions, 25-37, Interferometers, 835-837
76 Log-periodic, 767-769, 848
Thin wire antenna, 556 Long wire, 821
Thinned arrays, 68-73, 364 Loops, 747, 751,762
Non-isotropic elements, 72 Monopoles, 734-748, 783, 789,
Probabilistic studies, 69-72 790,792,797-805,807,
Space tapering, 72-73 814,819,841,847,849,
Trailing wire antenna, 715 854,855
Transmission lines Multiple installations, 850-856
Coaxial, 860-871 Navigation systems, 775
Conical, 887-901 Notches, 739,745-747,762,
HF, 722 804,807,855
Impedance calculation, 902-905 Omnidirectional antennas, 726-
Open-wire, 871-880 753
Strip lines, 883-886 Point to point links, 775
Transmission-line antennas, 584-586, Position-finding antennas, 830-
737,739-743,800,810 839, (see Position finding
Travelling wave antennas antennas)
Arrays, 161 Reflectors, 754-763
LF antennas, 634 Slots, 737, 743-745,749, 751,
Linear array feeds, 80, 87-98 761-762,765,772-773,
Long-wire, 677-678 803,809,816,820,841,
Patch arrays, 130-133 848-850
Vee antenna, 132 Transmission line antennas, 737,
Travelling-wave tubes, 98 739-743,800,810
Turnstile antennas, 748, 751-752, Turnstile, 748, 751-752, 821,
821,847 847
Vehicular antennas, 776-830,
UHF antennas (see VHF and UHF (see Vehicular antennas)
antennas) Whips, 790, 792, 793, 800
Umbrella antenna, 588-596,684 Yagis, 760, 765-767, 796, 810,
Uniphase antenna, 556 816, 848
Unipole (folded), 604 VHF instrument landing array, 757
VLF ground loss resistance, 572
VHF and UHF antennas, 725-859 VLF, LF and MF antennas, 553-662
Adcock array, 831-834 Active antenna, 639-640
Alford loop, 747, 847 Aircraft antennas, 6 25
Applications, 725-726 Base-loaded T-type radiator, 578
Backfire, 758-759 Broadcast antennas, 606-624,
Circularly-polarised antennas, (see Broadcast antennas)
845-850 Conductor losses, 577
Design parameters, 775 Diamond antenna, 586
Dielectric covers, 842-844 Dipole antennas, 625-627
Dielectric-loaded antennas, 840- Dual MF frequency feeding, 652-
844, (see Dielectric loaded 655
antennas) Earth conductivity, 574-577
944 Index

Feeding methods, 578-579 Vertical antennas, 554-577, (see


Ferrite loop, 637 Vertical antennas)
Folded monopoles, 604-606 Vertical monopoles, 598-603
Folded umbrella, 596 Wire antennas, 625
Franklin antenna, 556,603, 610 Yagi antenna, 655
Frequency range, 553 Vee antennas, 679-680,697
Ground conductivity effects, Vehicular antennas
569-574 Adaptive arrays, 416,417
Ground loss resistance, 553 Aircraft (see Aircraft antennas)
Ground systems, 569-574 Bicone, 793
Grounded antenna, 554 Cars and vans, 788-791
Half-delta loop, 636 Circular arrays, 321
Half-wave radiator, 556 Coaxial dipoles, 792, 793
Helix inductors, 577 Conducting surfaces (influence
Insulators for masts, 643 of), 781-787
Inverted L-type radiator, 578 Dipole parallel to cylinder, 786
Inverted-V dipole, 634 Dipoles on ground, 625-627
Large VLF antenna, 587-588 Environmental constraints, 787
Litz cables, 577 Ground effects, 633
Loaded antennas, 578-598 HF antennas, 711-717
Loop,637 Helically loaded antenna, 629
Masts (base insulators), 640-643 Impedance of, 782-784
Masts (guyed), 640 Interaction with platform, 6 29-
Mechanical design, 640-657 631
Mobile antennas, 624-633, (see Land vehicles, 711-713
Vehicular antennas) Land-based vehicles, 788-793
Monopole, 554 Long wires, 821
Multiple-tuned VLF antennas, Marine vehicles, 716-717, 793-
597-598 794
NATO VLF station, 655 Mathematical modelling, 777-
OMEGA antenna, 595 780
Portable antennas, 624-633 Measurements on vehicles, 827-
Propagation factor, 553 830
Receiving antennas, 636-640 Mechanical constraints, 787
T-type antenna, 580-584 Military vehicles, 791-793
Top-loaded antenna, 554, 560, Missiles, 819-820
578,584,589 Monopoles, 783, 789, 790, 792,
Top-loading mechanical design, 819
644-648 Nose spike, 820
Tower antenna, 561 Persons and animals (antennas
Tower heights, 553 on), 823-827
Transmission line radiator, 584- Propagation characteristics, 789
586 Radiation pattern, 785-787
Travelling wave antenna, 634 Ray tracing techniques, 779-780
Tuning coil losses, 577 Rockets, 819-820
Tuning units, 64 8-6 5 2 Satellites, 821-823
US Navy Northwest Cape VLF Scale modelling, 781
antenna, 587 Ships, 716-717
Umbrella antenna, 588-596 Siting methods, 776
Umbrella type top-loaded radi- Slots, 820
ator, 579 Tuning electrically short anten-
Uniphase antenna, 556 nas, 631-633
Unloaded antennas, 598-606 Turnstile, 821
Index 945

VLF, LF and MF antennas, 624- Waveguides


633 Bifurcated T, 98
Whips, 627-629,631-633, 790, Brick waveguide array, 215
792,793,823 Circular radiators, 284-288
Wire-grid modelling, 778 Corporate feed, 161
Vertical antennas, 554-577 Distributed waveguide array, 169
Calculation of parameters, 557- Fourier transforms, 286-288
559 Linear waveguide array, 154
Current distribution, 557 Multimode array elements, 123-
Current element, 555 126
Electrical height, 564-565 Periodically loaded, 98
Finite conductivity effects, 574- Rectangular radiators, 284-288
577 Slots, 113-121,142, 183
Ground conductivity, 569-574 Whip antennas, 627-629, 631-633,
Ground system, 569-574 687,711,715,716,790,
Grounded monopole, 554, 569- 792,800,823
574 Wire antennas, 181
Impedance, 562-564 Wire grid array, 164
Induced EMF analysis method, Wire grid lens, 709-711
558,561 Wire grid modelling, 713, 717, 778
Method of images, 559 Woodward and Lawson synthesis, 43
Moment method of analysis, 559 Wullenweber arrays, 299, 301, 319,
Monopole, 554 324,697-699
Mutual impedance between
monopoles, 565-569 Yagi antenna (Quater MF station),
Optimum height, 556 655
Performance parameters, 556- Yagi-Uda array, 765-767, 816
557 Arrays of, 760, 765,810
Radiation resistance, 559-561 Bandwidth, 766
Top-loaded antenna, 554, 560 Crossed, 848
Tower antennas, 561 Dielectric loaded, 841

Water at HF, 672


THE HANDBOOK OF ANTENNA DESIGN

Editors: A.W. Rudge, K. Milne, A.D. Olver and P. Knight

l  Authored by a multi-national group of antenna experts of inter­national standing.


l Presents the principles and applications of antenna design, with emphasis
upon key developments in the last 15 years.
l Fundamental background theory and analytical techniques explained in detail
where appropriate.
l Includes extensive design data and numerous examples of practical application.
l Deals with a very wide range of antenna types, operating from very low
frequencies to millimetre waves.
l New measurement techniques described in detail.
l Covers associated topics such as radomes, array signal processing and coaxial
components.
l Includes design data for antennas for satellite and terrestrial com­munications,
radar, mobile communications and broadcasting.

About the Editors

Dr Alan Rudge has had more than 15 years experience in antenna research
and design at Universities and Institutions in the U.K. and the U.S.A. He is currently
Managing Director of ERA Technology Ltd in the U.K., where he has established an
international centre for antenna research and design.
Dr Ken Milne is an independent electronic systems design consultant. He was
previously Director of Research for Plessey Radar Ltd, where he specialised in Radar
Systems and Microwave Antenna Design.
Dr David Olver is at Queen Mary College, University of London, where he
lectures and researches in microwaves and antennas. He is Honorary Editor of IEE
Proceedings ‘Microwaves, Optics and Antennas’,
Dr Philip Knight is with the British Broadcasting Corporation (BBC) Research
Department where he has worked on broadcast antennas and propagation problems
associated with broadcasting.

Peter Peregrinus Ltd


Southgate House, Stevenage, Herts,
SG1 1HQ England
ISBN 0 906048 82 6
Printed in the United Kingdom

You might also like