Understanding Radar Systems
Understanding Radar Systems
Understanding Radar Systems
UNDERST ANDING
RADAR SYSTEMS
Simon Kingsley
DeparlftJenl of Electronic ond Electriea/ Engineering
University of Sheffield
Shaun Quegan
Deparlmelfl of Applied and Campu.lalional Mathematics
UniwrsitJ' of Sheffield
10 9876543
ISBN 1-891121-05-7
ISBN 13: 978-1-891121-05-0
SciTech Publishing
911 Paverstone Drive, Suite B
Raleigh, NC 27615
Phone: 919-847-2434
Fax: 919-847-2568
www.scitechpub.com
CONTENTS
Preface xi
Acknowledgements xiii
t Fundamentals t
1.1 What is radar? 1
1.2 A simple radar explained 3
1.3 Overview of radar frequencies 4
1.4 Antenna gain 6
1.5 The radar equation 11
1.6 Accuracy and resolution 15
1.7 Integration time and the doppler shift 18
1.8 Summary 22
1.9 References 23
1.10 Further reading 23
1.11 Problems 23
3 Tracking radar 48
3.1 Introduction 48
3.2 Sequential lobing 49
3.3 Conical scanning 49
3.4 Monopulse radar 50
3.5 Tracking accuracy 54
3.6 Frequency agility 55
3.7 The tracking process 56
3.8 Radar guidance 58
3.9 Summary 60
3.10 References 60
3.11 Further reading 60
3.12 Problems 60
4 Radar detection theory 62
4.1 Introduction 62
4.2 The basis for decision making-probability theory 66
4.3 The effects of the receiver on the noise distribution 73
4.4 The distribution of signal plus noise 78
4.5 The signal-to-noise ratio 79
4.6 Detection and false-alarm probabilities 81
4.7 The correlation receiver 84
4.8 The matched filter 90
4.9 Key elements of signal detection 91
4.10 Detection using multiple observations 91
4.11 Modifications for intermediate-frequency input
to the receiver 94
4.12 Target fluctuations-the Sweriing cases 95
4.13 Summary 99
4.14 References 101
4.15 Problems 101
5 Signal and data processing 104
5.1 Introduction 104
5.2 Properties of clutter 106
5.3 Moving-target indicator processing 107
5.4 Fast Fourier transform processing 112
5.5 Thresholding 115
5.6 Plot extraction 118
5.7 Plot - track association 119
5.8 Track initiation 120
5.9 Tracking 122
5.10 Summary 125
5.11 References 126
5.12 Problems 127
CONTENTS vii
Appendices 326
I Symbols, their meaning and SI units 326
II Acronyms and abbreviations 330
III Useful conversion factors 333
IV Using decibels 335
V Solutions to problems 337
Bibliography 364
Index 367
PREFACE
Understanding Radar Systems is a book to convey facts and figures and also
explain why things are the way they are. It is written for students and young
engineers in industry, already competent in electronic engineering or physics,
who need to understand modern radar principles, applications and some of
the jargon used.
There are three parts to the book. The first three chapters form an easy
(and deterministic) guide to the basics of what radar is all about, and what
it can do. Chapters 4-6 are possibly more difficult because probability is
introduced, but these chapters will be necessary for those who wish to begin
calculating the performance of radar systems. The remaining chapters
describe different types of modern radar, the problems and the research areas.
These applications are used to introduce some more new ideas, so that their
relevance is apparent.
This book can be used in several ways: as a 'streetwise' introduction
to radar; as a rough guide to calculating the performance of most types of
radar; or as a handy reference when starting off on a new radar topic, before
going on to the more detailed texts in that area. The most important
thing is to understand the subject and enjoy your career in radar.
ONE
FUNDAMENTALS
• What is radar?
• What can it do?
• How well can it do it?
Radar is all about using radio waves to detect the presence of objects and
to find their position. The word radar, first used by the US Navy in 1940,
is derived from radio detection and ranging, thus conveying these two purposes
of detection and location. Modern radar goes further and is being developed
to classify or identify targets, and even to produce images of objects, for
example mapping the ground from a satellite.
The principle of radar is that a transmitter sends out a radio signal,
which will scatter off anything that it encounters (land, sea, ships, aircraft),
and a small amount of the energy is scattered back to a radio receiver, which
is usually, but not always, located near the transmitter. After amplification
in the receiver, the signals are processed to sort out the required echoes from
the 'clutter' of unwanted echoes by a combination of both electronic signal
processing and computer software (data processing).
There are many applications for radar. on scale sizes that vary from a
few centimetres, such as the measurement of the thickness of furnace walls,
to long-range systems probing planets across the solar system. Table 1.1
2 UNDERSTANDING RADAR SYSTEMS
Civil
Ground-based Air traffic control
Sea traffic control
Weather forecasting
Speed traps
Intruder alarms
Radar astronomy
Ground probing
Industrial measurement
Sea-borne Navigation
Collision avoidance
Air-borne Altimeters
Navigation
Weather
Space-borne Studying Earth resources
Sea sensing
Manipulating spacecraft
Mapping planets and minor bodies
Military
Detection of own forces or enemy forces
Tracking of air, sea, land or space targets
Guidance of own weapons systems
gives some idea of the variety of applications, and the list is growing as radar
systems find their way into industry, homes and even the motor car.
There are many systems similar to radar, such as sonar (which uses
sound instead of radio waves), medical ultrasonics and passive detection
systems t. Knowledge of radar principles is a good starting point for
understanding these other subjects and would help a physicist or electronic
engineer, for example, to transfer into medical electronics and body scanning.
Just as in the movie industry, where the apparently immense task of
making a film can be broken down into identifying 60 or 70 key scenes, so
other problems can be broken down. If you find the prospect of learning all
about radar a bit daunting, remember that there are less than 100 key things
you really need to know to understand the basics of radar, and these are
summarized for you at the end of each chapter, as they occur.
t An example of a passive (no transmitter) system is the technique for locating thunderstorms
using several receivers to obtain bearings on the radio signals emitted by lightning flashes; the
lightning is, in effect, acting as the transmitter for the rest of the system. Biologists tracking
animals tagged with tiny radio transmitters have a similar interest in passive detection and
location.
FUNDAMENTALS 3
Figure 1.1 shows a basic radar system. For this example we have chosen a
pulse radar, although we shall be looking at other forms of radio transmission
later. The principal radio frequency (RF) of the radar, the carrier, is set by
the frequency synthesizer. This continuous signal is pulsed on and off (usually
spending much more time off than on) by the modulator. The short bursts
of radio energy that result are amplified by the transmitter and sent to the
antenna via a switch called variously a transmit -receive switch, a T - R switch
or a duplexer. There are different designs for these switches, but they all
have the same two functions: to connect the antenna to either the transmitter
or the receiver at the appropriate times, and to protect the receiver from the
full force of the transmitted pulse.
When a pulse is transmitted, the radar clock begins to count time. The
radio pulse travels away from the radar at the speed of light, is scattered
from a target and returns to the radar. The distance to the target, R (called
the range of the target), can then be calculated from the time delay.
Remembering that velocity = distance/time, we can rearrange this as
[m] (I.l )
'Baseband'
RF
(e.g. S-band)L-------'
mixer
I
/' "'
{ Raw \
,
\ displa\ I
'- ./
/
Figure 1.1 Block diagram of an elementary radar system, and some abbreviations commonly used.
4 UNDERSTANDING RADAR SYSTEMS
Radio waves are part of the electromagnetic spectrum and so they could be
described by either quantum theory or wave theory. It turns out that, at the
frequencies used by radar, it is much more useful to think in terms of waves,
and we do not refer to quanta in the rest of this book. The frequency f
[Hz = s - 1] t of an electromagnetic wave is related to the wavelength A. by
C = fA [m s - 1 ] ( 1.2)
t The hertz [Hz] is named after Heinrich Rudolf Hertz, a professor at Karlsruhe, who is
widely credited as having founded radar because of his work in the late 1880s, which included
the reflection of radio waves generated by a spark gap generator.
FUNDAMENTALS 5
A B C D E FGHI
, J K LM
NATO system
2 - N ' 'T\Ooco 0 0 - xGHz
I/)
N _ "T~
- N
::;;
Figure 1.2 Radar frequency band names. The code letters L, S, C, X and K were used for
security reasons during World War II and have been widely adopted by radar engineers ever
since, despite the introduction of more rational systems.
The antenna is the device for radiating and receiving electromagnetic energy
and it has three main functions:
2. To provide for beam steering such that some area of coverage can be
provided.
3. To permit the measurement of angular information so that the direction
of a target can be determined.
When considering radar antennas it turns out that much of the theory
developed in physics courses for optics and optical astronomy is very useful.
The concept of reciprocity is especially helpful in allowing us to say that, if
transmitter power is fed into an antenna and it spreads out in a certain
angular pattern, then, when the same antenna is used for receiving, it has a
sensitivity to incoming signals that follows the same pattern. It is common
practice to mix these two cases, and you can hear antenna patterns
and polar diagrams being discussed one moment as if the antenna were
transmitting and the next moment as if it were receiving.
If an antenna were omnidirectional, it would radiate power uniformly
over the whole 4n steradians of a sphere. (Do not worry if you have not
come across steradians. They are the unit of 'solid' or three-dimensional
angle, and are not crucial to what follows.) Such an antenna is called an
isotropic radiator, but it would not be useful to us because, first, it would
not help with the three functions we require of it and, secondly, it can be
demonstrated mathematically that it cannot exist in practice. It is more
common to build antennas designed to focus or concentrate power in a
certain direction. This property of beaming power is known as the directive
gain D(O, <jJ) or the power gain G(O, <jJ). These terms are often abbreviated
to directivity D or gain G.
While the directive gain describes the radiation pattern of the antenna,
it does not give any indication of how lossy the antenna is. The power gain
includes the concept of losses, which occur through heating of the antenna
itself, the ground plane and any matching devices, as well as power radiated
into sidelobes. Power gain may be defined as the ratio of the radiation
intensity in the main lobe of the antenna to the radiation intensity from a
100 per cent efficient isotropic antenna having the same power input. The
gain of a lossless antenna is given by
[ ] ( l.3a)
where 110 = width of the beam in the azimuth direction [radians] and
11<jJ = width of the beam in the elevation direction [radians].
The azimuth angle 0 gives directional or bearing information and, as on
a magnetic compass, it is measured clockwise from the north. The elevation
angle <jJ is measured from the horizon upwards. Any point in the sky can be
specified by quoting its azimuth and elevation (see Fig. 1.3). If 110 and 11<jJ
8 UNDERSTANDING RADAR SYSTEMS
/ G a i n (6. d»
--~
Horizontal
Relative gain
Elevation d>
used to denote the maximum gain of an antenna, but beware of the frequent
use of the term to describe how the antenna's radiation pattern changes
with angle-'the gain falls off rapidly with elevation angle ... ', etc.
The maximum gain of an antenna can also be calculated from its size:
[ ] (t.5 )
where Ae = effective area of the antenna. The effective area is usually less
than the real area A (area of a parabolic dish, for example) and is related
to it by
Ae = eA (1.6 )
Here e [ ] is an efficiency factor, usually falling in the range 0.4-0.9 for
parabolic dishes, which arises because it is difficult to illuminate a dish
perfectly from the radio source at the centre. Some antennas, such as those
used by domestic televisions, can have effective areas greater than the physical
cross-section of the device.
Where does Eq. (1.5) come from? Any source of waves having a linear
dimension d, which is large compared to the radio wavelength, tends to have
a characteristic beam width A() in the appropriate plane given by
[radians] (1.7 )
We can explain Eq. (1.7) using an optics analogy. Wavelets coming from
many small (Huygens) sources along the aperture d interfere at some distant
point, as in Fig. 1.4a. Directly in front of the source, we find that the
wavefronts add constructively and a beam is formed; but as we begin to
move off-axis, they begin to interfere with each other because of the differing
path lengths. The actual signal received at a point off-axis depends on the
(a) (b)
+-------d-------+
Figure 1.4 (a) Wavelets from small Huygens sources add coherently in front of the aperture.
(b) At an angle sin () = A./ d they interfere destructively and the antenna gain falls to zero, thus
giving an approximate characteristic beamwidth to the first nulls of 22/d and A./d to the -3 dB
(half-power) points.
10 UNDERSTANDING RADAR SYSTEMS
---
Figure 1.5 The azimuth and elevation beamwidths arising from a rectangular aperture.
FUNDAMENTALS 11
On the return path this power again spreads out over the sphere of area
4nR2. Although it does not usually spread out uniformly, the 'gain' of the
target is automatically included in the concept of the ReS. The power density
at the radar thus becomes
(1.11)
but we will adopt the convention that Ls is always less than 1, and therefore
appears on the top. Using this definition, the power received by the radar,
from the target, is given (with appropriate units also shown) by
__ Pt [W]Gt [ ]Gr [ ]a [m 2 ]).2 [m 2 ]Ls [ ]
P [W] (1.13)
r (41r)3 [ ]R4 [m4]
[ ] (1.l4 )
This is the all-important radar equation, which is much used in one form
or another. It is perhaps a little too complicated to learn, and the best
strategy is probably to remember how to derive it quickly from the basics.
Often, the radar equation is used to solve for one unknown. For example,
supposing a particular SNR is required for reliable target detection (a typical
figure might be 13 dB, which is 20 times noise). The maximum detection
range Rmax of a given radar can be calculated from
[m] (1.15)
In the example above, the two really big factors are the R4 propagation
losses and the noise. Very little can be done about either of these and the
overall detection range of radars is thus fairly predictable; it is the cleverness
and the adaptability of radars that improves with time, rather than the
absolute detection range. Also, the radar cross-section used in Eq. (1.14)
fluctuates continuously for most real targets and we must use statistical
means to describe it. This infers that the radar equation itself is a statistical
method of detecting targets and measuring range, rather than being a
deterministic calculation; we say more on this in Chapter 4.
How well can a radar measure range? To answer this, we must be a bit more
careful with our question and define what we mean by the words 'how well'.
The range accuracy indicates the uncertainty in a measurement of the absolute
distance to an object, whereas the range resolution tells us how far apart two
targets have to be before we can see that there are indeed two targets rather
than one large one.
The range resolution of a radar system is fairly straightforward. If the
time delay between the echoes from two objects is greater than the pulse
duration r, then two separate echoes are seen (see Fig. 1.6a). If the targets
are closer than r, the echoes will merge (Fig. 1.6b). When the echoes are
separated in time by an amount similar to the pulse duration, then they
become resolvable (Fig. 1.6c; see also Fig. 6.1). This is like the Rayleigh
criterion used in optics, and combining it with Eq. (1.1 ) gives an expression
for the range resolution llR:
llR = crf2 [m] (1.16)
Usually the radar receiving system samples the receiver output every
r seconds, and each sample represents a distance llR called a 'range gate'
or a 'range bin'. A radar using a pulse duration r = 1 JlS would thus employ
16 UNDERSTANDING RADAR SYSTEMS
(a)
Target I Target :1
(0)
(e)
Figure 1.6 Radar range resolution. (a) Two targets are easily resolved when they are more
than a pulse length apart, (b) unresolvable when they are much closer than a pulse length, and
(c) just resolvable when separated by a pulse length.
a 1 MHz AID converter to sample the receiver output every 150 m in range,
out to the practical maximum range of the system.
Unfortunately, understanding the range accuracy of a system is not as
simple as the resolution. Intuition tells us that the accuracy of a range
measurement should depend on the 'sharpness' of the pulseshape, and it is
therefore rather surprising to discover that the crucial factor determining the
range accuracy is the bandwidth occupied by the radar. Figure 1.7 gives some
insight into why there is this dependence on bandwidth. The system shown
in Fig. 1.7a emits a single continuous tone and measures the phase of the
echo to find a rough position for the target, but it is ambiguous every
wavelength. In Fig. 1.7b a second frequency is added to the transmission to
reduce ambiguities and sharpen the position of the target. Adding further
frequencies (more bandwidth) eliminates the ambiguities and gives even
greater accuracy to the range measurement (Fig. 1. 7c).
In practice, pulseshape and bandwidth are related in simple pulse radars.
Short pulses take up more bandwidth B of the radio spectrum than long
pulses, a result well known to those familiar with Fourier transforms and
FUNDAMENTALS 17
E
(b)
(e)
E
Target
Figure 1.7 Radar range accuracy dependence on bandwidth. (a) With a single frequency,
accuracy is poor and the target position is ambiguous. (b) With two frequencies, accuracy and
ambiguity are improved. (c) Using more bandwidth, ambiguities are removed and accuracy
improves further.
Imagine that you are out walking and, glancing briefly over your shoulder,
you spot a helicopter flying in the distance. With such a short glimpse it is
FUNDAMENTALS 19
impossible to tell whether the helicopter is moving or not, and you stop to
look more carefully. It soon becomes obvious that the helicopter is not
moving quickly, but is it hovering or moving slowly? You stare at it for a
long time and decide that it really is hovering. What you have discovered
is that measurements of the position of a target can be made quickly, but it
takes time to estimate velocities and to distinguish differences in velocities.
The smaller the velocity difference, the longer the time needed to estimate it.
The length of time taken to make an observation with a radar set is
called the integration time, because all the data on a target are integrated or
added up until the measurements are sufficiently accurate. It is exactly like
photographing moving objects: a night-time photograph of the sky showing
a meteor burning up, taken with an exposure of 1/1 OOOth of a second,
will show only a faint object, which appears almost frozen in the sky;
another photograph taken by leaving the shutter open for half a second will
show a much brighter object, whose speed can be estimated from the length
of the trail.
How can the speed of an object be measured by radar? There are two
methods. The simpler method is similar to watching the helicopter or taking
a photograph, in that the speed of a target is estimated from the way its
position changes with time (except that optical systems measure transverse
velocities and position, whereas radar range information gives the radial
component). Observing changes in range is not a very accurate method for
a radar to measure speed, although tracking the target for a long time can
improve the estimate. This technique is used by incoherent radar systems, in
which the receiver is tuned to the same frequency as the transmitter but is
not phase-locked to it, and so is unaware of any small drifts in frequency
between them.
A more accurate way of measuring target speed is to make use of the
doppler shift, which is the change in the frequency of the radio signal caused
by the motion of the target. The doppler t shift is named after Christian
Johann Doppler (1803-1853), who pointed out that the colour of a luminous
body and the pitch of a sounding body are changed by the relative motions
of the body and the observer. In order to detect small changes in frequency,
a coherent system is needed in which the transmitter and receiver oscillators
are phase-locked to reveal any difference in the echo frequency.
Figure 1.8 shows a radar observing a jump-jet. If the jet is hovering,
each radio signal sent out by the radar will return with the same phase (as
measured with respect to the transmitted signal). If the jump-jet approaches
the radar, then each radio signal has to travel a shorter distance and the
phase of the echo will change continuously with time. Every A/2 through
t Although it seems disrespectful to previous generations of scientists, the modern convention
is that capital letters are not used to describe the effects named after them (e.g. gaussian noise,
doppler shift). The exception is when their name is used as a unit, for example electrical currents
are measured in amperes (or amps), but the unit is A.
20 UNDERSTANDING RADAR SYSTEMS
Figure 1.8 The motion of a target causes a change of phase in the radar signal, equivalent to
a frequency shift.
which the target moves means that the path length has been shortened by
A. and the phase of the echo will have changed by 2n or, in other words, will
have rotated by one complete cycle. If the jet were to fly at ;./2 [m s - 1 ]
towards the receiver, the radar would detect a 1 Hz change in the radio
frequency. If the jet flies at Vr [m s -1], then we must work out what this
represents in multiples of A.12 in order to calculate the corresponding doppler
shift Id; this leads directly to the formula (another one worth learning)
[Hz] (1.20 )
where Vr = radial component of the target speed towards the radar.
Transverse components of velocity do not contribute to the doppler shift
because the target neither approaches the radar nor recedes from it. Because
radars measure only radial components and optical systems, such as television
and infrared cameras, measure only transverse information, it is not
uncommon to find the two types of system integrated together in 'point' or
'local' defence arrangements.
The expression in Eq. (1.20) is actually an approximation to the full
relativistic formula, but it is valid for all cases where Vr « c, which is almost
always the case-very few radar targets approach the speed of light!
Very roughly, the integration time t needed to resolve two doppler
frequencies separated by Aid is given by
[s] ( 1.21 )
and using Eq. (1.18) again we can say that the error bid in the measurement
of doppler frequency is given by
Combining this with Eq. (1.21) means that the error c5vr in the velocity
measurement is
A.
c5v ~ --,---- (1.23 )
r 2tJ(2 x SNR)
The formulae above are not exact, because it all depends on how the
measurements are made. For example, if the doppler measurement is made
during a single long pulse, then the exact formula depends on the shape of
the pulse; if the measurements are made by integrating many pulses, then
the accuracy of the result depends on the 'weighting' or how the information
is put together. More details can be found in Chapter 4 and in Skolnik I.
This frequency must be distinguished from the land echoes, for which
Id = 0, and so the frequency difference !lId = 0.1 Hz. From Eq. ( 1.21 ) we
can say that the observations must be made over to s for the ship to be
distinguished. A pulsed HF radar usually sends out several hundred
pulses per second, and so a few thousand echo pulses would be collected
together for doppler processing when the radar was in ship detection
mode.
The microwave radar uses A. = 0.1 m, and when this is inserted in
Eq. (1.20) the doppler shift for the ship comes out to be 100Hz. An
observation period of only 0.01 s or to ms would therefore be needed
to separate the ship and land echoes, and typically this would be achieved
using a short burst of about 16 pulses.
A SNR of 20 dB means that the signal power is 100 times the noise
power, and putting this into Eq. (1.23) gives velocity errors of 0.5/fi == 0.35
ms- I for both radar systems. Try it!
22 UNDERSTANDING RADAR SYSTEMS
1.8 SUMMARY
Radar is used to detect the presence of an object and measure its position
and speed. It is possible to make a swift estimate of the performance of a
radar system using the formulae given below.
Key equations
[ ]
• Range resolution:
AR = c,/2 [m]
• Bandwidth:
B", II, [Hz]
• Range error:
c
bR ~ 2BJ(2 x SNR) [m]
• Doppler shift:
fd = 2vr l A [Hz]
FUNDAMENTALS 23
<>/. 1 [Hz]
d - tJ(2 x SNR)
• Error in measuring radial velocity:
A
<>v ----,--------
r 2tJ(2 x SNR)
1.9 REFERENCES
1. Introduction to Radar Systems, M. I. Skolnik, McGraw-Hill, New York, 1985. [This is the
classic paperback on radar systems, but is more of a reference work than a good read.]
2. Introduction to Optics, F. L. and L. S. Pedrotti, Prentice-Hall, Englewood Cliffs, NJ, 1987.
3. The Plessey Type 46C Weather Radar, RSL 1479, Issue 8, 1983.
4. Racal-Decca 2690 BT Marine Radar Series, Racal Marine Electronics publication reference
RMEjOOO8j1187 JAD.
5. Plessey ACR 430 Airfield Control Radar, RSL, Issue 2,1987.
1.11 PROBLEMS
1.1 The Racal-Decca 2690 BT Marine Radar Series4 offers radar options that include (a) a
2.7 m wide antenna operating at a wavelength of 3 cm and (b) a 3.6 m antenna operating at
10 cm. Calculate the horizontal beam width of the two antennas, assuming each has a linear
efficiency factor of 80 per cent.
24 UNDERSTANDING RADAR SYSTEMS
1.2 The vertical beamwidth (-3 dB) is quoted as 23" for the 3 cm version of the 2690 BT
system. What would you expect the antenna gain to be?
1.3 One of the 2690 BT radar bandwidth settings is 4 MHz. What range resolution would this
give?
1.4 The Siemens- Plessey ACR 430 5 is an X-band airfield control radar for the close control
of aircraft approach to land in poor weather conditions. The effective aperture size is 3.4 m
horizontally by 0.75 m vertically. What would be the beamwidths at an operating frequency of
9.40Hz?
1.5 What antenna gain would you expect for the ACR 430 antenna?
1.6 During evaluation of the ACR 430, aircraft were positioned for approach by controllers
using the radar for the first time. Averaged over 41 approaches to a runway, the worst
mean offset from the centreline approach was less than 25 m at a range of 4 n. mile (see
Appendix III). How good is this performance, and would it be sufficient to position an aircraft
actually on a runway in poor visibility?
1.7 Compare the detection ranges of the two 3 cm radars above for a 10 m 2 target using a
single pulse. Assume both radars have a mean noise level of -'131 dB W, have losses of - 5 dB
and require a SNR of 13 dB for reliable detection. The marine radar has a peak transmitter
powerof25 kW, but the airfield control radar is more powerful with a peak power of 55 kW.
CHAPTER
TWO
DESIGNING A SURVEILLANCE RADAR
• Surveillance
• Choosing parameters
• Radar cross-sections of targets and clutter
• Final design
Marine radar, air traffic control, air-borne radar defence systems, ground-
based radar arrays searching for satellites in space-these are all well known
examples of radar surveillance systems that work in all weather conditions,
and at all times of day, testifying to the versatility of the radar technique.
Most modern surveillance radars are 'multifunction', jargon for their
capacity to carry out other activities at the same time as searching for new
targets. One of these functions is to keep track of existing targets until they
are within a certain distance of the radar, when an entirely separate system
takes over: at a civil airport, for example, a Terminal Area radar guides in
the planes; a ship under attack, on the other hand, might engage a close
support weapons radar to lock on to the target. In this chapter we are going
to concentrate on one of the main applications of surveillance radar, which
is to search the sky continuously for new targets.
2S
26 UNDERSTANDING RADAR SYSTEMS
Surveying the sky presents the radar design engineer with something of a
dilemma; the radar needs to scan the whole sky and get back to the starting
point as quickly as possible in order not to miss any new developments.
However, the radar also needs to spend as much time as possible staring at
each part of the sky in order to obtain good results. Perhaps the best way
to illustrate the difficulties is for us to follow a typical design process and
develop the problem for ourselves.
A common choice of frequency for long-range surveillance is L-band
( '" 1.3 GHz) because this avoids the bad-weather problems that can affect
higher frequencies. S-band ('" 3 GHz), for example, is often used only for
medium-range surveillance up to about 60 nautical miles or 111 km. In
practice, many factors would determine the final choice of frequency,
including the type of target to be detected and the coverage required. We
will assume that a frequency of 1300 MHz ( = 1.3 GHz) has been chosen,
When considering the antenna design, there are two conflicting require-
ments. First, there are several good reasons for choosing narrow beams:
1. The angular position of the target can be measured with good precision.
2. The number of unwanted echoes cluttering up the picture at anyone
time is reduced.
3. The number of interfering signals that can get into the beam at anyone
time is also reduced.
4. The antenna gain factors Gt and Gr in the radar equation are increased
(i.e. the transmitter power is more concentrated in the direction of the
target) and, consequently, the signal-to-noise ratio is improved, making
the target easier to detect.
This leads to the general rule that an antenna with a gain of G must probe
DESIGNING A SURVEILLANCE RADAR 27
SOLUTION The physical area of the dish is 28.3 m 2 but, because of the
efficiency factor, this reduces to an effective area of 17 m 2 • The wavelength
at 1300 MHz is 23.1 cm and thus the gain of the antenna is given by
G = 4n:Ae/ A.2 ~ 4000 []
There are therefore 2000 beam positions in the sky to be searched.
The width of the radar beam is given roughly by Eq. (1.7) as
A 0.231 .
110 = 11</> '" - = - - =: 0.039 radians =: 2.2°
d 6
but again effidency considerations mean that in practice the dish would
not be uniformly illuminated by the source antenna at the focus and the
effective beam would be broader, perhaps 3° wide.
COMMENT Two thousand beam positions may seem a lot to search, but
a 3° beamwidth is wide by today's standards, and radars with narrower
beams have even more positions to search.
The greater the rate at which pulses are transmitted by a radar system, the
greater the mean power radiated, and it is interesting to ask the question:
'In the design of a radar system, what is the constraint on the maximum
rate at which pulses can be transmitted?' The limit occurs when pulses are
transmitted so frequently that one pulse is transmitted before the previous
pulse has completed the round trip to the target and back. In this situation,
it is unclear which transmitted pulse originated which echo pulse, and the
target range becomes ambiguous.
The rate at which pulses are transmitted is called the pulse repetition
frequency, often abbreviated to PRF. Using Eq. (1.1) and the' 150 m per
microsecond' rule we can say that, if the time interval T between infinitesimally
short transmitter pulses were 1 J1S (a PRF of 1 MHz), then ranges up to a
maximum of 150 m could be surveyed before the receiver had to be turned
off to allow the next transmitter pulse to be sent. More practically, for a
PRF of 1 kHz, range ambiguities are 150 km apart (although the range that
can be surveyed is a little less than this because of the pulse duration and
T - R switching times). The maximum PRF that can be used for an
unambiguous range Rmax is given by
PRF ~ c /2Rmax [Hz] (2.3 )
What would happen if we tried to survey ranges up to 150 km with a
1.5 kHz PRF (which has an unambiguous range of only 100 km)?The answer
is shown in Fig. 2.1, in which a continuous sequence of transmitter pulses
and the associated echoes from a target are shown. The most obvious
assumption is that echo pulses 1 and 2 come from the transmitted pulses A
and B respectively, meaning that the target is within the 'first-time-around'
range interval of 0-100 km. In the case shown in Fig. 2.1, the first-time-
around range is 30 km. However, it is also possible that echo 2 came from
Unambiguous range
of 100 km UOkm
I' l
i i i i i i
Transmitter Echo Transmitter Echo Transmitter Echo
pulse pulse pulse pulse pulse
A I B ;
Figure 2.1 A radar system with an unambiguous 'first-time-around' range of 100 km. The
target range appears to be 30 km, but it could be 130 km.
DESIGNING A SURVEILLANCE RADAR 29
transmitter pulse A and echo 1 came from the previous transmission. In this
case, the target lies in the' second-time-around ' range interval of 100-200 km
and the target range is 30 + 100 = 130 km. Likewise it is possible to have
third-time-around echoes in the range 200 to 300 km, and so on.
SOLUTION The maximum PRF can be found directly from Eq. (2.3) as
333 pulses/second. Another way to look at the problem is to rearrange
Eq. (1.1) as
T= 2R/c [s]
which tells us that when R = 450 X 103 m the round-trip time for a radar
signal is T = 3 ms. For a pulse emission every 3 ms, the PRF must be
l/T = 333 pulses/second, often referred to as a PRF of 333 Hz.
The duty cycle is the ratio of the pulse length r to the inter-pulse
period T. In this case
r 3 x 10- 6
Duty cycle = - = == 10- 3 []
T 3 X 10- 3
COMMENT Often, this PRF is too low to adequately sample the doppler
frequency. Also, the duty cycle may be too low for certain types of
transmitters that need to be worked harder than this if sufficient power
is to be developed to give a good performance over 450 km.
There are several ways in which the PRF can be increased without
reducing the unambiguous range. One method is to 'colour' or label each
pulse in some way so that it can be distinguished from its neighbours.
Methods of labelling pulses include transmitting them with different fre-
quencies, phases, polarizations or pulseshapes. There are difficulties with
these methods, however, when the target radar cross-section (RCS) fluctuates
from pulse to pulse. Labelling pulses also interferes with doppler processing
and moving-target indication (MTI)-see Chapter 5-and so has only
limited usefulness.
Another method of increasing the PRF is to use bursts of pulses on
different PRFs. Each one of these independently may be ambiguous over
the range interval, but when used in combination with the others the
ambiguities can be eliminated. Two staggered PRFs are demonstrated in
Fig. 2.2; when PRF (a) is used the target range is measured as 30, 180, or
330 km, but when PRF (b) is used the range is measured as 105 or 330 km.
The true range of the target must therefore be 330 km.
30 UNDERSTANDING RADAR SYSTEMS
(a)
Target!
~~V/N'''-''''''''_~......J!(,-___~____~_. ~
(0) () 105 225 330 km
Figure 2.2 The two PRFs (a) and (b) give ambiguous range information when used
independently. In combination, only one range is possible.
In this section we wish to show that short transmitter pulses imply fast
sampling of the receiver output, which can create difficulties for the digital
signal processing. Even with the recent rapid growth in computing power
we must still be careful not to design a radar that produces digits so quickly
that we cannot afford to buy computers large enough to process them on-line.
The pulse length of 3 p.s chosen above is fairly typical for a long-range
air surveillance radar, and from Eq. (1.16) we know this gives a nominal
range resolution of about 450 m. In practice, the resolution would be worse
than this because of pulse shaping and other losses in the systems, and the
final figure might be nearer 750 m. On the other hand, Eq. (1.19) tells us
that the range can be measured more accurately than the nominal 450 m if
we have a good signal-to-noise ratio. These improvements in the estimates
of range come from two processes. The first is known as 'plot extraction'
and involves interpolating between adjacent range samples on each scan.
The second process is tracking, which is the smoothing of the apparent path
of the target through many observations. The overall range accuracy of the
system would be about 50 m after all these processes had been undertaken,
and this would normally be adequate for long-range air surveillance.
The output of our radar receiver should be sampled every 3 p.s because,
as Fig. 2.3a shows, the samples are then separated by a time/distance equal
to the transmitted pulse length (as measured at the 3 dB or half-power
points). Because the echo pulses must be at least the same length as the
transmitted pulses, this sampling rate ensures that no information can be
missed. If the sampling were carried out less frequently than every 3 p.s,
information would be lost and a small target might escape detection by
falling between two samples, as shown in Fig. 2.3b. There are some benefits
in sampling more frequently than every 3 p.s (known as 'oversampling')
because it improves the SNR on a target that straddles two range bins
(Fig. 2.3c ) but it involves a lot more signal processing for only a small reward.
By 'sampling' we usually mean using an analogue-to-digital (AID)
converter to represent the voltages coming out of the receiver by binary
numbers that digital and computer hardware can process. If a sample must
be taken every 3 p.s, this implies an AID converter speed of 333 kHz.
Radar systems need a high 'dynamic range', meaning that they must be
able to process large echoes from nearby objects (often clutter) at the same
time as echoes from distant objects that are very faint; this is a natural
consequence of the 1I R4 term in the radar equation. The bigger this range
of voltages in the receiver, the greater the number of bits needed in the A/D
converter to represent them and it is roughly true to say that one more bit
is needed every time the voltage range is doubled ( = 6 dB increase in dynamic
range). However, the dynamic range requirement can be eased in practice
by using range-dependent gain or swept gain. During the transmission of each
32 UNDERSTANDING RADAR SYSTEMS
Transmitter Echo
pulse pulse
(a) Power
Target
t t t t t AID samples
Target detected
(b) Power
t t t t AID samples
Target missed
(c) power, ( \
l/ ~~_TimeOrrange
t t t t t t iii t t t t AID samples
Target detected
SNR improved
I I I I I I I I • Range Iml
0 450 900 1350 1800 2250 2700 3150
I
0
I
3
I
6
I
9
I
12 15
I I
18
I
21
. .
Time ljJ.s]
Figure 2.3 (a) The receiver output sampled correctly; (b) undersampled, which may cause a
target to be missed; and (c) oversampled, improving the SNR for a target straddling two range
bins.
pulse, the receiver is held at zero gain; at short ranges the gain is gradually
increased and at long ranges the full receiver gain is used. Beyond 40-50 km,
clutter decreases owing to the curvature ofthe Earth and shadowing effects.
In our surveillance radar we will assume that an 8-bit word is insufficient
for each A/D sample and that 2 bytes per sample must be used.
Our receiver output is now being sampled at 666000 bytes per second,
and worse is to come; we need two receiver channels in order to measure
the doppler information. A single coherent receiver can reveal the speed of a
target but it cannot tell whether it is moving towards, or away from, the
radar. A second receiving channel is used to resolve this ambiguity by shifting
it 90° in phase from the first channel; these are known as I and Q channels,
which stands for In-phase and Quadrature. This second channel must also
DESIGNING A SURVEILLANCE RADAR 33
be sampled at 666000 bytes per second and the overall data rate produced
by our radar is thus about 1.3 Mbytes per second. This sampling rate is not
too great for modern computers to store or move around in memory, but
it becomes more challenging if complex floating-point calculations have to
be carried out. Short-range radars working with much higher bandwidths
and shorter pulse lengths than our surveillance system can present the design
engineer with some tough and expensive data processing problems to solve.
I. For target sizes »)., the ReS is roughly the same size as the real area
of the target. This is known as the optical region because the ReS
approaches the optical value.
2. For target sizes ~ }., the ReS varies wildly with changes in wavelength,
and it may be greater or smaller than the optical value. This is known
as the resonance or Mie region.
3. For target sizes <d, the ReS ex ). -4. This is known as the Rayleigh
region after Lord Rayleigh, who discovered that the scattering of light
by particles in the atmosphere varies as ). -4 (Tyndall, Mie and Oebye
also contributed to these studies). This wavelength-dependent scattering
explains why the sky is blue and the sun appears yellow. When white
light from the sun arrives at the Earth, the relatively long-wavelength
yellow Ired components of the spectrum pass more or less straight
through the atmosphere compared with the shorter-wavelength blue
34 UNDERSTANDING RADAR SYSTEMS
light, which is scattered over the sky. Much the same is true on a larger
scale size where low-frequency radar signals are undisturbed by water
droplets in rain and clouds, but millimetric radar signals suffer significant
scattering. This reasoning lay behind our choice of the relatively
low-frequency L-band for surveillance at the beginning of the chapter.
The ReS of a few simple shapes are given in Table 2.1, for the case
when the object size is large compared with a wavelength. These simple ReS
values turn out to be quite useful for several reasons. First, it is sometimes
possible to get a feeling for the Res of an object by building it up out
of a few simple shapes; for example, see the chapter by J.D. Olin in
reference 1. Secondly, at long wavelengths, targets often behave as uncompli-
cated structures because the scattering is not from many tiny surfaces but
Table 2.1 The ReS of some simple shapes when the size of the object is
large compared with a wavelength
Sphere
G
+
More general large
curved surface ~/r2
\
'(I /
/ \
\
/
\ I \ /
\ I \ I
'./ \,1
3
16n r4 2 (Jd4nrSin ()/).))2
--cos ()
).2 4nr sin ()/).
Circular flat plate
broadside
4
4nr 2 (sin(2nrSin()/).))2
-cos2
()
). 2nr sin () / ).
Square flat plate
broadside
DESIGNING A SURVEILLANCE RADAR 35
Circular cone
Dihedral corner
reflector
r
• ~,
Trihedral corner
reflector
'I \t 121tr4
'2
I.
~2
Chaff
N
-
,,12
-0.lSNi.2
This can be a useful way of thinking about RCS when the target is an
antenna, for which the concept of gain is well defined.
When the target is not simple, and cannot be synthesized from simple
shapes, there are various well established computer programs that can be
applied to the problem of calculating the RCS. If all else fails, a scaled copper
model of the target can be built and measured in an anechoic chamber at
an appropriate scaled frequency. A good summary of the state of the art was
presented in 1989 when an entire issue of the Proceedings of the IEEE was
devoted to the radar cross-sections of complex objects2.
In the case of our L-band radar, the radar cross-section of an aircraft
will lie in the optical region because the wavelength of 23 cm is much less
than the size of the target. We cannot assign a single value to the RCS
because it will depend on the aspect angle at which the target is viewed,
both in azimuth and in elevation, and also on the polarization angle of the
radar. These factors, combined with interference from different scattering
surfaces on the target, mean that as we observe the aircraft the RCS will
fluctuate. These fluctuations can be treated by finding the mean value of the
RCS (Jay and a probability density function (PDF) p«J) to describe the
variations about the mean. A well known density function for random
variables is the chi-squared variable (X 2 ) with two, or four, degrees of
freedom. The forms are
4(J
p«(j) = -Zexp (2(j)
-- [ ] (2.7)
(jav (jay
In 1960, P. Swerling 3 used these expressions as the basis for four proposed
mathematical models describing different types of RCS fluctuation. These
four Swerling cases are discussed in Chapter 4 when we examine target
DESIGNING A SURVEILLANCE RADAR 37
detection theory, and they are still in widespread use today, even though
more sophisticated models have been developed since. With low-frequency
systems, such as over-the-horizon radar, the less complicated scattering
mechanisms mean that quite often the RCS of a target remains constant for
a considerable period of time; such non-fluctuating targets are sometimes
referred to as Swerling case 5 targets.
The time-averaged (RMS) value of the RCS (Jay is sometimes used in
the radar equation because this is the only information available. However,
the fluctuations mean that the predicted performance will only be achieved
for that part of the time when (J ~ (Jay' The Swerling models predict RCS
values lower than average to occur more frequently than above-average
values because the amplitudes and phases of two interfering signals vary
independently (see Chapter 4). Amplitude and phases must both be nearly
equal for above-average values of (J to occur. The PDF for the first case is
shown in Fig. 2.4, and it can be seen that values of (J significantly above (Jay
rarely occur. The use of (Jay in the radar equation thus gives an over-optimistic
performance estimate, and it becomes important to know which model best
describes the RCS fluctuations of the target. Appropriate adjustments can
then be made when calculating the probability with which a target can be
detected by a particular radar system, or what coherent gain may be assumed
when pulses are added together to improve the signal-to-noise ratio.
For our surveillance radar, typical values of RCS that might be expected
are given in Table 2.2. Further information may be found in Maffett 7 •
1.0
Probability density function
0.5
Figure 2.4 The probability density function for a Swerling case I target.
38 UNDERSTANDING RADAR SYSTEMS
RCSon RCS on
linear log
Target scale scale
2.6 CLUTTER
Clutter is a single word used to describe all the unwanted echoes that clutter
up the radar picture. Clutter is nearly always present and usually widespread,
with echoes arising from hills, buildings, the sea, birds and insects, meteors,
the aurora and many other sources. Of course, what is clutter in one
application may not be so in another. An example of this is HF radar, which
is often used to observe waves on the sea, in which case echoes from ships
(ship clutter) confuse the picture; but when HF radar is used for over-the-
horizon ship tracking, it is the echoes from waves (sea clutter) that cause
the problems.
Clutter may occur as distributed clutter, which increases with the
resolution cell size, or as point clutter, which does not. Point clutter arises
from discrete scatterers such as electricity pylons. The term surface clutter
is used to describe land or sea echoes from the area illuminated by the radar.
The radar cross-section of the clutter is best described by calculating the
average RCS density <To, the RCS per unit area, which is given by the ratio
[ ] (2.8 )
where <Tc = RCS of the area Ac. The symbol <To is sometimes referred to as
sigma zero, and it can be thought of as representing the radar reflectivity of
the terrain.
DESIGNING A SURVEILLANCE RADAR 39
RMI1R
Radar
Figure 2.5 The area of ground illuminated by a radar beam at grazing incidence.
40 UNDERSTANDING RADAR SYSTEMS
(provided the radar was sited high enough to see the ground at this range)
would be roughly
Area = 50 x 103 x 450 x 0.05 == 1.2 x 10 6 [m 2 ]
A reflectivity of 1/100 suggests that the clutter from the ground would be
12000 m 2 • Assuming the radar is searching for an aircraft of RCS 1 m 2 , the
signal-to-clutter ratio would be 1/12000. In decibels, this figure corresponds
to 41 dB, and is one of the factors that determine the dynamic range
requirement for the receiver. In practice, at least another 13 dB of dynamic
range is needed, so that an adequate SNR is available for target detection,
plus some extra allowance for clutter variations around the mean value.
COMMENT With increasing range, the signal strength from a target decays
as R4 because of the radar equation but the clutter power falls off as R3
because the area illuminated has a linear dependence on R. Signal-to-
clutter problems therefore get worse at longer ranges.
If target echoes are so much smaller than clutter, how can they be
detected at all? The answer is that most targets of interest are moving and
can be distinguished from stationary clutter through the use of the doppler
effect. Doppler processing to give sub-clutter visibility (SCV) is an important
part of radar processing and is discussed in Chapter 5. However, the
DESIGNING A SURVEILLANCE RADAR 41
Power [dB]
proportional to RCS [dB m C]
+50
13 d-B-f----f"---;.rn'g' f'";rem".
-20 -10 0 10 / 20
Cruise missile Doppler shift [Hz)1
implication for our surveillance radar design is that we must dwell for long
enough in each beam position to gain sufficient doppler resolution to separate
target and clutter echoes.
Most surveillance radars also maintain clutter maps, which can be static
(loaded prior to operation) or dynamically modified during operation.
Separate maps are maintained for ground clutter, rain, chaff, etc. The radar
antenna beam and the signal processing can then be adaptively modified to
suit the environment in a process known as clutter fixing.
2.7 NOISE
Before we can proceed with our final design, we need to know what noise
level to expect. At L-band the noise is likely to be dominated by internal
noise produced by the random motion of thermally excited electrons. Below
about 6000 GHz (i.e. for all practical radars), thermal noise can be considered
to be white noise having a flat power spectral density S(f) given by
S(f) = kTo (2.12 )
where k = Boltzmann's constant = 1.38 x 10- 23 J K -1 and To = system
temperature [K]. To is generally assumed to be 290 K, giving a noise density
of - 204 dB W Hz - 1, a figure engraved on the hearts of most radar engineers.
This noise spectrum extends far beyond the bandwidths of the radar system
and we are only concerned with band-limited white noise, which has been
restricted by the passband of the receiver. The mean noise power in the
42 UNDERSTANDING RADAR SYSTEMS
receiver N is given by
[W] (2.13 )
where B = bandwidth [Hz]. Note that this figure is independent of the radar
operating frequency; an S-band and an L-band radar, both operating with
a bandwidth of 1 MHz, would have identical mean noise levels of -144 dB W
(= -204 dB W Hz- 1 + 60 dB Hz bandwidth).
In practical receivers, the noise level is found to be worse than in
Eq. (2.13 ) by a factor known as the noise figure F. The noise figure is measured
over the linear part of the receiver operating range as
F = noise power out of actual receiver
(2.14 )
noise power out of ideal receiver
In the case of our L-band surveillance radar, the bandwidth can be found
from Ij(pulse length) = 333.3 kHz = 55.2 dB Hz. Assuming a noise figure
for the receiver of 3 dB, the mean noise level is
N = kToFB [dBW]
-204 + 3 + 55.2 [dBW] (2.15 )
= -145.8 dB W
Tx filter Beamformer
Isolator Rx protection circuit
Connectors Rx front-end filter
Radome (protective antenna coating)
The instantaneous noise power in the receiver will fluctuate about this mean
value. In Chapter 4 we discuss the statistical properties of noise in more
detail, and the way in which this affects the probability of target detection.
Having decided most system parameters, we can now work out the
transmitter power needed to undertake the task of surveillance. We will
assume that the targets will be no smaller than 1 m 2. Rewriting the radar
equation (Eq. (1.14» gives
(SNR)N (4n)3 R~ax
Pt = - - - - - - : : - - - - [W] (2.16 )
Gt Gr}·2 a L s
Next we insert the values chosen during this chapter, i.e.
Pt 16 [MW] 72 [dB W]
p= ~ [W] (2.18 )
t (PRF)t
The terms Band t are collected together because their product is approxi-
mately unity. This 'pulse radar equation' is easy to convert to an energy
equation by replacing the SNR ratio with the energy Inoise ratio (E IN) and
the transmitted power ~/(PRF) by the transmitted energy Et • Thinking of
the radar equation in terms of energy can be a useful concept when there is
a fixed time interval for data collection, as is the case with our surveillance
radar.
In our final design, we would probably integrate 32 pulses in each beam
position for the following reasons:
All these reasons dictate the need for an observation time of nearly
100 ms in each beam position for successful target detection. At the beginning
of the chapter we worked out that only 5 ms per beam position is available
if the whole sky is to be scanned. This reveals one of the truisms of
surveillance-there is never enough time to do everything properly and
compromises have to be made. There are several ways out of the dilemma;
perhaps the most obvious method is to use several beams simultaneously,
DESIGNING A SURVEILLANCE RADAR 45
for this reduces the rate at which a dish must be scanned. An option available
to the most modern radars, where the beam can be steered electronical1y
(Chapter 15), is to make intelligent decisions about how the search pattern
should be adapted to the situation, according to a set of priorities. In some
cases, the simplest solution is simply to abandon searching certain parts of
the sky, usual\y those at high elevation angles.
In real-life radar design there are many more sophistications and
problems to be dealt with. There would be integration losses during the
32 pulses because ofRCS fluctuations, a variety of modulation schemes would
be used, and so on. But these are all complications that will be easily mastered
if you understand the basic design process and the physics of what is
happening in radar survei11ance.
2.9 SUMMARY
Radar surveillance can be improved through the use of narrow beams, but
this may lead to there being more beam positions to be searched than is
possible in the time available. Multibeam systems can help with this problem,
but they put more pressure on the data processing activities, which are often
already stretched, even with today's technology.
The RCS of real targets fluctuates and its statistical nature must be taken
into consideration if the radar detection performance is not to be over-
estimated. The problem of clutter and the need for sub-clutter visibility is
often severe and leads to a need for doppler processing. Careful design
of the transmitted waveform is needed to avoid range and/or doppler
ambiguities.
Key equations
• Number of beam positions to fill the sky (hemisphere):
Number of beam positions = G /2 []
• Maximum PRF for an unambiguous range Rmax:
PRF ~ c/2R max
• Radar cross-section of a target:
a = 4nA 2 / i. 2 [ m2 ]
• Radar cross-section per unit area:
aO=ac!Ac []
• Radar cross-section per unit volume:
'1=ac/~
46 UNDERSTANDING RADAR SYSTEMS
2.10 REFERENCES
2.11 PROBLEMS
2.1 The Italian company Alenia manufactures the ATCR-44K, an L-band medium-range
primary radar designed for use in modern automated air traffic control (ATC) systems. The
antenna is a rotating reflector and if a rotation rate of 6 revolutions per minute (RPM) is
selected then typically a pulse repetition frequency (PRF) of 480 Hz and a pulse duration of
1.5 JlS is used. Alternatively, at 12 RPM the PRF is typically 691 Hz and the pulse duration is
1 JlS. Given that the peak transmitter power is 1.2 MW, what is the duty cycle and the mean
power for the two modes of operation?
DESIGNING A SURVEILLANCE RADAR 47
2.2 Derive the expression, given by Alenia, that the number of hits n on a target each time the
antenna scans past is given by
M(PRF)
n=
(RPM) x 6
where !l(} = azimuth beamwidth [degrees]. If !l(} = 1.2°, how many hits per scan are there
when the antenna rotation rate is 6 RPM?
2.3 The documentation for the Alenia ATCR-44K radar gives the maximum unambiguous
range as
1000000
R =----- [nautical miles]
max 12.4 X (PRF)
What do you think the number 12.4 represents?
2.4 What is the detection range of the ATCR-44K radar for a 2 m 2 target? Assume 5 dB system
losses, a further 2 dB for atmospheric absorption, and that a 10 dB SNR is required. Take the
elevation beamwidth as 4.7".
2.5 Does the ATCR-44K radar require staggered PRFs?
2.6 What AjD converter requirement might the ATCR-44K radar be expected to have?
2.7 During World War II the German Navy used a 125 MHz medium-range early-warning
radar known as Freya, The rotatable antenna consisted of 12 dipoles mounted in front of a
rectangular wire mesh reflector (see Swords 8 ). If the peak transmitter power was 20 kW and
the pulse duration 2 ps, at what range could the radar detect a 2 m 2 target? Assume an antenna
gain of 16 dB and losses of 5 dB. How does this compare with the modern Alenia radar?
CHAPTER
THREE
TRACKING RADAR
• Measuring angle
• Monopulse radar
• Tracking accuracy
• Radar guidance systems
Tracking radars are dedicated to a target and are designed to measure angular
information accurately.
3.1 INTRODUCTION
48
TRACKING RADAR 49
G
I
-----
--
"
'-
'-
..... ,
\
I
Two heams
...
"0
.E
Q..
E
'-
' ..... _--/
I
<
Sum ~
Difference .l
Figure 3.2 Monopulse antenna beam patterns expressed (a) in polar form and (b) in
rectangular coordinates.
52 UNDERSTANDING RADAR SYSTEMS
the antenna to get back on target. It should be stressed that the presence
of the phase-sensitive detector does not mean that the system exploits the
phase information contained within the radio echo as a means of deriving
angular information.
Amplitude monopulse can be improved by taking the ratio M which
normalizes the difference channel by the sum channel. This gives a quotient
that is independent of the signal strength and linear against the angle error
over a wide range of angles. The function of dividing signals can be under-
taken digitally, but in the past it has been performed by processing the signals
with logarithmic amplifiers and then taking differences.
Perhaps the best way to get to grips with amplitude monopulse is to
draw out the antenna patterns in Fig. 3.1 for yourself; try using a pocket
calculator to generate a (sin e/ e) 2 pattern, plot two versions of it with one
slightly displaced from the other, and then add them together to form the
sum channel and subtract them to form the difference channel. Chapter 7
of Hirsch and Grove 3 describes in a particularly helpful way how to use a
computer to carry out a more sophisticated simulation of monopulse antenna
patterns and differences.
Phase-comparison monopulse is also possible and was tested early on in
radar history; in fact, the technique was patented in 1943 (see Rhodes 2 ).
Two antennas are fixed adjacent and parallel to each other and, by comparing
the phase difference of the two outputs, it is possible to derive angular
information. An echo arriving along the bore-sight of the antennas will arrive
at both of them at the same time (Fig. 3.3a), but a signal arriving at an
angle e to the bore-sight will arrive at one antenna later than at the other
because it has had to travel an extra distance x given by
x = d sin e [m] (3.1 )
where d = separation of the antennas Em]. The distance x can be expressed
as a fraction of the radar wavelength A to give the difference in phase 111/1
between the two signals as
111/1 = 2nd(sin e)/A [radians] (3.2 )
The factor 2n in Eq. (3.2) arises because the phase difference increases by
2n radians for every complete wavelength A. travelled by the signal. Note
that for small angles sin e "" e, leading to the approximation
111/1 "" 2nde / A [radians] (3.3 )
which is a roughly linear relationship between the angular deviation from
the bore-sight direction and the phase-difference error signal.
Phase-comparison monopulse is identical to the Young's slit experiment
of physical optics and the interferometry techniques used by radio astronomers.
However, the usefulness of phase-comparison methods in tracking radar is
somewhat limited because of problems created by multiple signals from
TRACKING RADAR 53
(a)
t tE
'.~oU ~----1.---- Bore-sight
(0)
Figure 3.3 Phase-comparison monopulse. (a) When the echo arrives along the bore-sight,
there is no phase difference between the signals. (b) When the echo arrives at an angle 0, this
gives rise to a phase difference 1'11/1.
At long range, a target can usually be considered as a point source and the
angular accuracy of a tracking radar is determined by both electromechanical
factors associated with the control of the antenna turning gear and by the
SNR of the radar measurements. Whichever tracking method is used, from
Eq. (1.18) we would expect the angular accuracy be to be related to the
beamwidth and to be fundamentally limited by
be,.... AeIJ(2 x SNR) [radians] (3.4)
However, the use of sum and difference channels allows this to be improved
upon such that be I k may be achievable, where k is the slope of the A/~
curve near e = 0, see Levanon 4 •
The use of sequential lobing, conical scanning or monopulse is merely
the means by which the precision of Eq. (3.4) is achieved, but note that the
first two techniques will be further degraded by amplitude fluctuations
whereas monopulse is not. In practice, RCS fluctuations, multipath and
changes in the atmospheric propagation all add together to increase the
tracking noise. These errors have been elegantly discussed and summarized
in Chapter 7 of Berkowitz s, for those who wish to delve deeper.
achievable in practice).
At a range of 10 km the transverse error R be corresponds to a
distance of 1.7 m. A 1 MHz bandwidth implies a basic range resolution
of 150 m, which would improve to 1.5 m for this SNR, so the two
accuracies are similar. It would be easier to improve the range accuracy
(using more bandwidth) than the angular accuracy, provided the
scattering size of the target was small enough to justify this.
At short ranges, the finite size of the target begins to limit the accuracy
of the system. Tracking radars need narrow 'pencil' beams and so tend to
operate at relatively short wavelengths, which puts the target RCS into the
optical region. At these short wavelengths the target behaves as many
independent scatterers, each of which contributes in a complex way (i.e. in
both amplitude and phase) to the overall RCS and causes an effect similar
TRACKING RADAR 55
The first task of a tracking radar is to acquire the target allocated for
engagement. Tracking radars usually have a search mode to survey a
restricted area of sky around the expected target position. Different search
patterns are used, but it is common to find options of a horizontal raster
scan, similar to a television, a vertical variant of this called a nodding scan,
and a low-elevation scan round the entire horizon, as shown in Fig. 3.4.
After the target has been detected, the next step is to estimate its position
in a process known as plot extraction. We have already seen how the angular
information is extracted, but on each pulse the target range and velocity
must also be measured.
A rough estimate of the range can be found by locating the range cell
in which the echo is largest, but much more precision can be obtained by
Nodding scan
~])'. '.
Figure 3.4 Various methods of scanning for a tracking radar to acquire a target.
TRACKING RADAR 57
interpolating between range cells. If two adjacent cells show equal echo
power, then clearly the target lies exactly half-way between them; but if one
is larger than the other, then some interpolation formula is required. Usually
the interpolation is carried out by fitting the pulse shape or a quadratic curve
to the data samples and then differentiating to find the peak. The error in us-
ing a simple quadratic rather than the actual pulse shape after matched filter-
ing is quite small. As an example, a quadratic curve can be fitted to the AID
samples shown in Fig. 3.5 using the equation Y =ax2 + bx + c. For the central
sample, x =0, so that c = yo; at x =+1 range cell, Y+ = a + b + yo and al- x =
-1 range cell Y- = a - b + yo. Solving these equations gives
The range of the target is then that of the central range cell + Xmax '
Amplitude [volts
or AID samples
representing volts]
L-------~~------~-r----L------L-- _________ _ + x
-1 +1
I
Xmax
(target position)
Figure 3.5 Interpolation using a cosine-squared pulse to improve the range accuracy.
58 UNDERSTANDING RADAR SYSTEMS
the missile in the centre of the radar beam. The attraction of this approach
is that it is simple, and a strong SNR is received by the missile because the
signal travels only on a one-way path. The disadvantages are that the target
must be kept accurately in the centre of the beam and, as with RCLOS, the
path up the beam may not be the ideal trajectory for a missile to follow 8 .
The disadvantages of beam riding can be overcome to some extent
through the use of semi-active homing in which the tracking radar acts as a
target illuminator. On board the missile a receiver detects the scattered energy
from the target, tracks its position and works out the best trajectory for an
intercept. The.missile does not have to fly within the illuminating beam and
the tracking radar is not required to keep the bore-sight exactly on the target,
merely to keep the target illuminated.
Although semi-active homing requires more complicated electronics to
be carried by missiles, it has proved to be an effective system. During the
Gulf War, for example, the Lynx helicopters of the UK Royal Navy
illuminated Iraqi ships with their radars and attacked successfully with their
semi-active homing Sea Skua missiles. The Patriot missiles, used with such
telling effect in the interception of Scud missiles during the same crisis, also
use semi-active homing, with the passive radar receivers on board the missiles
relaying information back to the main radar processor on the ground, which
then feeds commands back up a link to the missile.
The most advanced weapons employ active homing radar guidance
systems in which the entire tracking radar is carried on board the intercepting
missile. Initially the missile is locked on to a target by its host, but once
fired it has a high degree of autonomy to pursue its target. Developments
in miniaturized analogue and digital electronics, target image processing and
the evolution of 35 and 94 GHz radar seeker heads mean that some very
sophisticated radar missile guidance technology should be available soon.
The first of this new generation of missiles is likely to be the ERINTs
(extended-range interceptors), which use active homing and are so small
that 16 of them will fit into a launcher that at present holds four Patriots.
Worked example The new generation of anti-armour submunitions are
expected to be tiny missiles (0.1 m by 0.6 m long), which are fired in
clusters but which contain their own 94 GHz guidance radars for
independent targeting 8 • What diameter would an older X-band guided
missile have needed in order to obtain the same angular accuracy?
SOLUTION A 10 cm dish at 94 GHz has a nominal angular resolution
of A.ld = 1.8°, although, in practice, this would probably be a little
over 2°.
If, for convenience, we assume an X-band frequency of 9.4 GHz, we
can see immediately that the older-type missile would have needed a
diameter of the order of 1 m to have a resolution comparable with the
proposed new submunitions.
60 UNDERSTANDING RADAR SYSTEMS
3.9 SUMMARY
Key equation
• The decorrelation bandwidth formula:
Af~ c/(21) [Hz]
3.10 REFERENCES
1. Technical History of the Beginnings of RADAR, S. S. Swords, Peter Peregrinus for the lEE,
Stevenage, Herts, 1986.
2. Introduction to Monopulse, D. R. Rhodes, McGraw-Hili, New York, 1959.
3. Practical Simulation of Radar Antennas and Radomes, H. L. Hirsch and D. C. Grove, Artech
House, Norwood, MA, 1988. [A useful book, including software listings.]
4. Radar Principles, N. Levanon, Wiley, New York, 1988.
5. Modern Radar, Ed. R. S. Berkowitz, Wiley, New York, 1965.
6. Introduction to Radar Systems, M. I. Skolnik, McGraw-Hili, New York, 1985.
7. Radar Design Principles, F. E. Nathanson, McGraw-Hill, New York, 1969.
8. The evolution of radar guidance, D. A. Ramsay, GEe J. Res., 3(2), 92-103, 1985. [The
whole of this special issue is worth reading.]
3.12 PROBLEMS
3.1 The Siemens-Plessey WF3 Wind finder radar is an X-band tracking radar operating at
9375 MHz that automatically tracks a corner reflector attached to a meteorological balloon.
Conical scanning of a 2° beam at 1500 RPM is used and the PRF is 625 Hz; how many pulses
per scan are there?
TRACKING RADAR 61
3.2 The radar in problem 3.1 has a peak transmitter power of 55 kW, an antenna gain of
35 dB, a receiver bandwidth of 2 MHz and a noise figure of 10 dB. Assuming 6 dB losses and
given the performance indication that a reflector of RCS 120 m 2 can be tracked at a range of
100 km, estimate the angular accuracy of a standard I minute observation.
3.3 A weapon control radar system on a modern fighter aircraft operates at Ill-band using a
0.75 m x 0.5 m planar antenna. In dog-fight mode, the radar scans a 20° x 20° field ahead of
the aircraft at a medium PRF of 10 kHz. If three pulse bursts of 32 pulses are required to
confirm a target detection, what is the repeat time to search the field?
3.4 The Marconi Radar Systems ST802 is a lightweight monopulse naval tracking radar with
a 2.4° beamwidth. If a 30 dB SNR echo were received from an attacking missile, what tracking
accuracy would you expect? Would this be sufficiently accurate to bring the missile down by
radar-controlled gunfire?
3.5 One method of improving azimuth information is to use doppler beam sharpening (similar
to synthetic aperture radar described in Chapter II). Here the motion of an air-borne radar
can be used to subdivide the antenna beam because the relative velocity of ground clutter varies
across the beam. Try these calculations:
(a) For a frequency of 10 GHz (l/J-band), a beamwidth of 4° and an aircraft speed of
300 m s - 1, what is the doppler spreading of the clutter across the beam when the azimuth angle
is 45° away from the direction of motion and the look-down angle is 10°?
(b) If the doppler resolution of the processing is 10 Hz, to what azimuth resolution would
this correspond?
(c) Would doppler beam sharpening be effective looking directly ahead of the aircraft?
CHAPTER
FOUR
RADAR DETECTION THEORY
4.1 INTRODUCTION
62
RADAR DETECTION THEORY 63
u(t)
h(t) Decision
--'-------=:"--.c +
H(w) y(t) = I'(t) + m(t) maker
Receiver
c(t) lI(t)
value A represents the scaling of the signal due to propagation, losses and
the RCS of the scatterer (see Eq. (1.13». For simple hard scatterers, A will
be a constant, but in many cases the target RCS will fluctuate (see Sec. 4.12).
The value rd represents the delay introduced by the propagation of the signal
to and from the target, and provides the measure of range.
At this stage, the presence of a signal will normally be difficult to detect,
because the energy in the signal is very small compared to the noise power,
and to a large extent the voltage trace is dominated by high-frequency noise.
It is the task of the receiver to extract all possible information about the
presence of a target before any decision is made. In order to make progress
with analysing how it should be designed in order to do this, we need to
know some basic facts about linear systems. These are covered in any basic
text on communications or signal processing, such as Schwartz2 or Stremler 3 ,
and you may wish to consult one of these texts to remind yourself of their
derivation. For the moment, all we need to know is that the behaviour of
a linear system is completely described if we know the way it reacts to an
input signal that is a short, sharp spike (an impulse or delta function). This
is the impulse response function h (t), which tells us the response to a delta
function b (t) at time t = o. Since the response cannot occur before the
stimulus, h(t) is causal, i.e. t < 0 implies h(t) = O. For a general input signal
I (t), the output 9 (t) is given by a convolution operation
In the integral s is treated as a time variable. If f (t) were the sum of two
signals, i.e. 11 (t) + 12 (t), then we can see straight away by substituting into
the integral that the response of the system to the sum is the same as the
sum of the individual responses. This is what we mean by linearity; it is also
called the superposition property. The other vital property of convolution
can be obtained by replacing I (t) by I( t - r d ), a delayed version of I (t).
If you do this, you find that the response is also delayed by rd.
The relations between the input and output signals take on a more
attractive form when we take Fourier transforms. This has the remarkable
property of converting convolution into multiplication, so that Eq. (4.1)
becomes
G(w) = F(w)H(w) (4.2)
Here the use of lower- and upper-case letters indicates a Fourier transform
pair, i.e.
I(t)-F(w)
The receiver will be treated as a linear filter with impulse function h (t)
and system transfer function H (OJ), so that the output from the receiver,
y (t), can be written as
y(t) = v(t) + m(t) = Au*h(t - r d) + n*h(t) (4.3)
where v(t) and m(t) are the responses to the signal and noise inputs
respectively, i.e.
and
Since, in practice, both u (t) and h (t) are of finite duration and h (t) will be
causal, we can change the integration limits and write the signal term as
I
L__ _ \ L
I \ \
I \ \
I \
I \
( \
I \
w w
(a) (b)
Figure 4.2 The effects of a bandpass filter on target detection. In (a) the passband is too wide,
allowing excessive noise to affect detection. In (b) the passband is too narrow, and signal energy
is lost.
correlation receiver. These are derived in Secs. 4.7 and 4.8. (The equivalence
of optimizing information extraction and maximizing SNR is not as obvious
as it may appear; Woodward 4 provides a classic analysis of the problem.)
Both before and after matched filtering, the trace is at the intermediate
frequency (IF), i.e. it is a narrow-band signal centred on the IF carrier
frequency, and it is normal to perform envelope detection to remove the
carrier before making any decision. However, since all the essential ideas are
unaltered if we ignore the carrier frequency and assume that the input to
the filter is at baseband, we have elected to make this simplification. This
keeps the mathematical complexity to a minimum. For the reader who prefers
a more precise treatment, in Sec. 4.11 we indicate the modifications necessary
to deal with the case that occurs in practice.
The trace y (t) available to the decision maker might appear as shown in
Fig. 4.3, and at each instant (corresponding to each range), it is necessary
to decide whether the value of y(t) indicates the presence of a target. For
example, should the spikes at t1 and t2 in Fig. 4.3 be attributed to anything
other than noise? This involves a judgement as to whether this value was
more likely to have arisen from noise alone or from signal plus noise. The
basis for this decision must be a knowledge of the frequency with which
different values of y (t) will occur in the two circumstances.
A very useful way to describe how frequently y(t) takes different values
is through its probability density function (PDF) Py (y). This function is
defined by the property that, for small dy, values of y(t) in the range
y ~ y(t) ~ y + dy will occur with frequency given approximately by py(y) dy.
RADAR DETECTION THEORY 67
6
4
-2
-4t
-6
Hence, as ~y tends to 0,
p{a:::;y(t):::;b} = I b
py(y)dy (4.7)
(Read the left-hand-side of this equation as 'the probability that the measured
value y(t) lies between a and b'. The right-hand side is the area under the
graph of Py (y) between a and b.) Since Py (y) is measuring relative frequencies,
py(y) ~ 0 (4.8 )
and
(4.9)
The second condition simply states that, when a measurement y(t) is made,
the value obtained is certain to be a real number.
We have already seen the exponential PDF (Eq. (2.6) and Fig. 2.4). It
is easy to show that this PDF meets the two conditions. It occurs in a number
of important places in radar applications. Of these, we shall meet it as a
model for clutter in synthetic aperture radar images, and as a model for
targets with fluctuating ReS (Sec. 4.12).
Three particularly useful parameters that can be extracted from the PDF
are the mean p., the variance (12 and the standard deviation (1. The mean is
one type of representative value for the random function y (t). The other
two measure the spread of values taken by y (t). We shall see that they also
have very simple physical interpretations for electrical signals.
In order to explain how these parameters are found from the PDF, it
is easier initially to do this for a digital signal. Suppose that y( t) can take
only a finite set of values {Yl' Y2' .""' YM} (which in modern radars is normally
the case, since the signal will have passed through an A/D converter). In a
long series of measurements, Yi will occur ni times. Then the total number
68 UNDERSTANDING RADAR SYSTEMS
(4.10)
The average square deviation from the mean is known as the variance, and
is given by
Here we have used Eqs (4.9) and (4.10) to get from the second to the third
line. There are thus two ways to find the variance. The first is to subtract
the mean and then average the square differences. The second is to find the
average square value, then subtract the square of the mean. Either way gives
the same answer. The RMS deviation from the mean is the square root of
the variance and is known as the standard deviation (1. In these expressions
we are implicitly assuming that these parameters do not depend on the time
of measurement (the system is stable).
We need to take averages at several points in this chapter, and we do
not want to have to write the full integral with the PDF every time. Hence
we use the notation E[ ] (for expectation or expected value) to stand for
the average value of whatever appears between the square brackets. As
examples, we could write
Jl = E[y]
It is worth remembering that taking the expectation is a linear operation,
so that E[y + z] = E[y] + E[z] and E[cy] = cE[y] if c is a constant.
RADAR DETECTION THEORY 69
The mean, variance and standard deviation have very direct interpretations
in circuit terms. During the time T, y(t) will be expected to have values in
the range y ~ y(t) ~ y + Ay for approximately At = Tpy(y) Ay seconds.
Hence, if we regard the voltage y(t) as an input to a 1 ohm resistance, then
y(t) has a De value and a mean power given by
1
Devalue;:::; - ~> At = LYPy(y) Ay [V] (4.12 )
T
and
1
Power;:::; - Ly2 At = Ly2py(y) Ay [W] (4.13 )
T
Allowing Ay to tend to 0, the sums on the right-hand side become integrals,
70 UNDERSTANDING RADAR SYSTEMS
which are E[y(t)] and E[y2(t)] respectively. Hence we can see that
1
= ~ feo exp( _Z2 /2) dz
v (2n) 0
_ ~ [k exp( -z2/2)dz
V (2n)]0
= t- <1>(k) (4.16 )
To get from line 1 to line 2 in this derivation we made the substitution
z=(n-J.l)/u
RADAR DETECTION THEORY 71
0.5
IL
Figure 4.4 The gaussian PDF for mean J.I and standard deviation u; the shaded area
corresponds to the probability of observing a noise value k standard deviations above the mean.
Table 4.1 Values of the function (J) (k), which correspond to the area shown
in the diagram for a mean-zero gaussian variable with standard deviation 1.
Extreme values of (J) (k) corresponding to false-alarm probabilities of 10 - 5,
10- 6 ,10- 7 and 10- 8 are given in Table 4.1(b).
(a)Values of <I>(k)
K 0 2 3 4 5 6 7 8 9
0.0 0.0000 0.0040 0.0080 0.0120 0.0160 0.0199 0.0239 0.0279 0.0319 0.0359
0.1 0.0398 0.0438 0.0478 0.0517 0.0557 0.0596 0.0636 0.0675 0.0714 0.0754
0.2 0.0793 0.0832 0.0871 0.0910 0.0948 0.0987 0.1026 0.1064 0.1103 0.1141
0.3 0.1179 0.1217 0.1255 0.1293 0.1331 0.1368 0.1406 0.1443 0.1480 0.1517
0.4 0.1554 0.1591 0.1628 0.1664 0.1700 0.1736 0.1772 0.1808 0.1844 0.1879
0.5 0.1915 0.1950 0.1985 0.2019 0.2054 0.2088 0.2123 0.2157 0.2190 0.2224
0.6 0.2258 0.2291 0.2324 0.2357 0.2389 0.2422 0.2454 0.2486 0.2518 0.2549
0.7 0.2580 0.2612 0.2642 0.2673 0.2704 0.2734 0.2764 0.2794 0.2823 0.2852
0.8 0.2881 0.2910 0.2939 0.2967 0.2996 0.3023 0.3051 0.3078 0.3106 0.3133
0.9 0.3159 0.3186 0.3212 0.3238 0.3264 0.3289 0.3315 0.3340 0.3365 0.3389
1.0 0.3413 0.3438 0.3461 0.3485 0.3508 0.3531 0.3554 0.3577 0.3599 0.3621
1.1 0.3643 0.3665 0.3686 0.3708 0.3729 0.3749 0.3770 0.3790 0.3810 0.3830
1.2 0.3849 0.3869 0.3888 0.3907 0.3925 0.3944 0.3962 0.3980 0.3997 0.4015
1.3 0.4032 0.4049 0.4066 0.4082 0.4099 0.4115 0.4131 0.4147 0.4162 0.4177
1.4 0.4192 0.4207 0.4222 0.4236 0.4251 0.4265 0.4279 0.4292 0.4306 0.4319
1.5 0.4332 0.4345 0.4357 0.4370 0.4382 0.4394 0.4406 0.4418 0.4429 0.4441
1.6 0.4452 0.4463 0.4474 0.4484 0.4495 0.4505 0.4515 0.4525 0.4535 0.4545
1.7 0.4554 0.4564 0.4573 0.4582 0.4591 0.4599 0.4608 0.4616 0.4625 0.4633
1.8 0.4641 0.4649 0.4656 0.4664 0.4671 0.4678 0.4686 0.4693 0.4699 0.4706
1.9 0.4713 0.4719 0.4726 0.4732 0.4738 0.4744 0.4750 0.4756 0.4761 0.4767
2.0 0.4772 0.4778 0.4783 0.4788 0.4793 0.4798 0.4803 0.4808 0.4812 0.4817
2.1 0.4821 0.4826 0.4830 0.4834 0.4838 0.4842 0.4846 0.4850 0.4854 0.4857
2.2 0.4861 0.4864 0.4868 0.4871 0.4875 0.4878 0.4881 0.4884 0.4887 0.4890
2.3 0.4893 0.4896 0.4898 0.4901 0.4904 0.4906 0.4909 0.4911 0.4913 0.4916
2.4 0.4918 0.4920 0.4922 0.4925 0.4927 0.4929 0.4931 0.4932 0.4934 0.4936
2.5 0.4938 0.4940 0.4941 0.4943 0.4945 0.4946 0.4948 0.4949 0.4951 0.4952
2.6 0.4953 0.4955 0.4956 0.4957 0.4959 0.4960 0.4961 0.4962 0.4963 0.4964
2.7 0.4965 0.4966 0.4967 0.4968 0.4969 0.4970 0.4971 0.4972 0.4973 0.4974
2.8 0.4974 0.4975 0.4976 0.4977 0.4977 0.4978 0.4979 0.4979 0.4980 0.4981
2.9 0.4981 0.4982 0.4982 0.4983 0.4984 0.4984 0.4985 0.4985 0.4986 0.4986
RADAR DETECTION THEORY 73
K 0 2 3 4 5 6 7 8 9
3.0 0.4987 0.4987 0.4987 0.4988 0.4988 0.4989 0.4989 0.4989 0.4990 0.4990
3.1 0.4990 0.4991 0.4991 0.4991 0.4992 0.4992 0.4992 0.4992 0.4993 0.4993
3.2 0.4994 0.4993 0.4994 0.4994 0.4994 0.4994 0.4994 0.4995 0.4995 0.4995
3.3 0.4995 0.4995 0.4995 0.4996 0.4996 0.4996 0.4996 0.4996 0.4996 0.4997
3.4 0.4997 0.4997 0.4997 0.4997 0.4997 0.4997 0.4997 0.4997 0.4997 0.4998
3.5 0.4998 0.4998 0.4998 0.4998 0.4998 0.4998 0.4998 0.4998 0.4998 0.4998
3.6 0.4998 0.4998 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999
3.7 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999
3.8 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999
3.9 0.5000 0.5000 0.5000 0.5000 0.5000 0.5000 0.5000 0.5000 0.5000 0.5000
We have already remarked that the thermal noise n(t) entering the receiver
has a zero-mean gaussian PDF. This is not an adequate description of the
noise at the point where the decision is made, because at that point the noise
(and any signal component) has passed through the receiver. Hence we need
to consider the noise and signal-plus-noise distributions at the output of the
receiver. Again, we have to use results from linear system theory (see
references 2 and 3). Initially we deal with the case where we have noise n (t)
alone on input, giving a noise term m (t) at the output from the receiver.
Though for a general input distribution the calculation of the output
PDF can present great difficulties, the situation for gaussian inputs turns
out to be remarkably simple. In fact, gaussian inputs give rise to gaussian
outputs. Hence all we need to know to describe the output distribution are
the output mean value and standard deviation. The mean is easily taken
care of, since it is related to the mean of the input noise by
(4.17 )
t For consistency with most modern books on radar and communications, alJ the sections
of this book dealing with Fourier transform ideas use radian frequency w rather than cyclic
frequency f. By contrast, most engineers in their normal discourse work in Hz, and most of
our discussions do the same. Since f = w /2n, conversions are easy enough; translating between
the two is just part of the radar game (but still annoying).
RADAR DETECTION THEORY 75
to the total noise power from a narrow band I1w of frequencies centred on
w is well approximated by Sm (w) 11m/ (21t).
Let us first apply these ideas to the input noise. An important idealization
assumes that the input noise can fluctuate infinitely fast, so that knowing
the value at any instant of time t gives no information about the value of
the noise at any other time, no matter how close it is to t. In correlation
terms, this means that the ACF is a delta function,
(4.24 )
We can relate the constant N /2 to the noise power per unit bandwidth and
to the system temperature. To do this, note that in either case the total power
in the bandwidth B is
-
1 f27tH N
-dw=NB (4.26)
2n _ 27tH 2
Hence the noise power per unit bandwidth is given by the constant N. We
have already seen in Chapter 2 that the power in bandwidth B is kToB, where
k is Boltzmann's constant and To is the effective noise temperature of the
receiver. Hence
N =kTo (4.27)
Rm("C) = -
N f 2
7[B
4n - 21EB
IH(w)j2ejw< dw (4.30)
N f21tB
CT~ = - IH(wW dw (4.31 )
4n - 21EB
This has a very simple interpretation, since the energy (or integrated gain)
(a) 6
4
~
~
::
-4
-6
(b) 6
4
-6
Figure 4.5 The effect of filtering white noise: (a) shows white noise, while (b) shows the output
when this is passed through a filter with a cosine-squared impulse response function.
RADAR DETECTION THEORY 77
Some of the more general results of this section are needed elsewhere in
this book (especially Chapter 5), but if the input is white or band-limited
white zero-mean gaussian noise with noise power N W per unit bandwidth,
they can be summarized very simply. In fact, the output is also zero-mean
gaussian with variance O'~ = N Eh/2, so has PDF
1
Pm(m) =J 2 exp( _m 2 /20'~) (4.34 )
(2nO'm)
It is then easy to find the probability that the output noise m (t) exceeds any
threshold.
H(w)
I MHz
A- 4 •
-wo wo
I A
y y
J
t w
100 MHz
H(w)
2 MHz
A
i w
100 MHz
Suppose now that a signal component is present at the input to the receiver.
We know that if a delayed signal Au(t - 't d ) enters the receiver together
with additive noise n(t), then the output is given by (see Eq. (4.3»
y(t) = v(t) + m(t) (4.35 )
where v (t) and m (t) are the signal and noise terms in the output. Since at
the point of making a decision the user does not know 'td or, indeed, whether
there is a signal present at all, all that can be said is that, if there is a signal
present, the output distribution at time t will be gaussian, but with an
unknown mean v (t). Hence the decision maker is trying to separate, on the
RADAR DETECTION THEORY 79
I'
I \
I \ Signal plus
I \ noise
I \
I \
I \
I \
\
' .....
Figure 4.7 The PDFs of noise and signal plus noise at the output of the receiver.
PY(Y)=J(21tCT~)exp -
1 (y 2CT~
- V)2) (4.36)
often without any clear distinction being made between them. For a voltage
trace y(t) = f(t) + n(t) consisting of a signal component together with a
noise component, the first definition is the ratio of the energy in the signal
part (integrated over some time T) to the expected energy in the noise over
the same time,
SNR = energy in signal [ ] (4.37 )
energy in noise
For mean-zero noise of average power (12, the energy received in time Tis
T(12, so that this definition can be written
SNR = energy in signal power in signal
[ ] (4.38)
T power in noise (12
This definition is of most use at IF, so that the integration can be carried
out over many cycles of the carrier frequency, while the modulation remains
essentially unchanged. In this case, the received signal v(t) effectively has
the form V cos (wot), where Wo is the carrier frequency and V is a constant.
Hence the signal power is V 2 /2, giving SNR = V 2 /2(12. This is the SNR
calculated using the radar equation in the preceding chapters.
The second definition compares the instantaneous value ofthe signal with
the RMS noise power,
SNR' = IV(tW/(12 [ ] (4.39)
(some authors use Iv(t)I/(1). This definition is most appropriate when
discussing baseband signals. The two definitions are related, since, in practice,
the baseband signals will actually be the envelope of an IF signal. Over
intervals short compared with the rate of change of the envelope (i.e. much
shorter than 1/ B, where B is the bandwidth of the baseband modulation),
the IF signal can be written V cos (wot), where the baseband envelope has
instantaneous value V. Hence
[ ] (4.40 )
For consistency, throughout this book we adopt the first definition, which
is the one used in the radar equation. Care must therefore be taken in
comparing our results with those of other authors, particularly in the sections
on the correlation receiver and matched filter (Secs 4.7 and 4.8).
Notice that neither of these definitions is particularly useful to describe
the signal-to-noise ratio prior to the receiver, because at this stage we regard
the noise as having a very large bandwidth, and hence very large total power
(for white noise, this power would formally be considered infinite). It is only
RADAR DETECTION THEORY 81
after passing through the receiver ( or the noise bandwidth is reduced in some
way) that the SNR becomes meaningful, since the noise power is then simply
NEh/2. Figure 4.7 indicates that, as the SNR increases, the whole of the
signal plus noise PDF moves to the right. This implies that the chance of
making the right decision when the measured signal is thresholded also
increases.
A threshold decision may be the right way to separate signal from noise,
but how do we choose the threshold? For any threshold Y, four types of
events can occur. These are:
The subscripts indicate the PDFs of noise (n) and signal plus noise (sn).
Using the PDFs of Fig. 4.7, the areas defining PCa and Pd are shown in Fig. 4.8.
It is clear that, whatever the value of Y, increasing it will cause both PCa
and Pd to decrease. This illustrates that there is always a trade-off between
improving detections and reducing false alarms. In practice, most radars are
operated at a threshold level giving very low false-alarm rates (and hence
long times between false alarms), for reasons discussed in Chapter 5. In most
places in this book we have adopted a false-alarm probability of 10- 6 as
our working criterion. We now show how this fixes the threshold level, and
hence the reliability with which we can detect targets.
Let us calculate the probabilities of a correct detection and a false alarm
for a threshold set at k times the RMS noise value. The constant k is known
as the (multiplication) margin; when expressed in dBs it is additive. If there
82 UNDERSTANDING RADAR SYSTEMS
(a)
(b)
Signal
plus
noise
Figure 4.8 The areas in the PDFs of (a) noise and (b) signal plus noise corresponding to the
false-alarm and detection probabilities for a threshold Y.
SNR
Worked example What would be the false-alarm rate for a threshold set
at three times the noise power; what would the single-pulse detection
probability be if the SNR for a single pulse was 10 dB?
or
$(4.75 - J(2 x SNR» = 0.3
Hence our working criterion needs a SNR of 15.6 (11.94 dB) for reliable
detection. 'Reliable' here means Pd = 0.8.
Up to now, we have ignored the form of the signal term v(t) coming from
the receiver, and have simply examined the implications of thresholding the
combined signal plus noise. In practice, the shape of the signal coming from
the receiver is largely under the control of the system designer, and its
properties are chosen to meet such specified criteria as desirable time or
range sidelobe levels, resolution, etc. The most fundamental criterion, against
which all other criteria must be balanced, is the ability to detect targets,
which, as we have seen, is a function of the SNR. Hence a natural priority
is to design the receiver to maximize SNR. By Eq. (4.33), we know that the
noise power (Jm is dependent only on the gain of the receiver, not on the
shape of its impulse response function. Hence, for fixed gain, the best SNR
is obtained by maximizing the response to the signal term. This is achieved
by an elegant, conceptually simple and readily implemented processing
scheme, known as the correlation receiver, whose frequency-domain imple-
mentation is the matched filter.
The scheme relies on a fundamental result known as the Cauchy-Schwartz
inequality. This states that, given two functions f (s) and 9 (s) of finite energy,
r
then
lIb f(S)g*(S) dS I
2
~ If(sWds Ib Ig(sWds (4.47)
where
• h(t) is causal
• Eh = Ell
• (J; = NEIl/2
To find when the peak SNR will occur, substitute this expression for h(s)
in Eq. (4.51). Then
v(t)=A tT u(t-'rd-s)u(T-s)ds
= A tT u(t - 'rd - T+ s)u(s)ds (4.55 )
(4.56)
and
Then
AJ(2Eu IN) ~ 1.77
We already know that we must use a value for k of 4.75, which gives a
detection probability
Pd ~ t- $(2.98) ~ 0.001
Such very low single-pulse detection rates are not uncommon for radars
using mUltiple-pulse detection techniques. These are discussed in Sec. 4.10.
There we show that, for coherent processing, using M pulses gives
a SNR improvement by a factor M. In this example, that means that
using 10 pulses would bring the detection rate up to 80 per cent.
The operation being carried out by the receiver is perhaps most clearly
explained by making the change of variables s = t' - t + T in Eq. (4.55).
This gives
v(t) = A II
I-T
u(t' - Ld)U(t' - t + T) dt' (4.60 )
Input signal
f-----'-::----~------ - - - - - - - --+
f--------L.,.----'------- - - - - - -
Output signal
f - - - - - - - - - - - - - ' - - = - - - - - ' - - - - - - -+ T d+ T
t - T
Figure 4.10 The response of the correlation receiver to a returned signal. The top plot shows
the return (delayed by 1'd) and the plots below show a succession of positions of the correlating
function. On the right is shown the complete time response of the receiver.
The signal term in the output from the correlation receiver is given by
(see Eq. (4.55»
which is the correlation of the signal with itself. This is known as auto-
correlation, and appears so often in radar processing that it deserves its own
notation. Since in general we need to consider complex signals f (t), we
define the deterministic autocorrelation function of a signal f (t) as
(a)
-ol
-6
(b)
-4
-6
(e)
-4
-6
30
(d)
25
20
15
Figure 4.11 The response of the correlation receiver to a signal contaminated with white noise:
(a) shows the signal (delayed by 'd); (b) shows the noise-alone trace; (c) is their summation;
and (d) is the squared response of a receiver matched to the signal shape of (a) when the
signal-plus-noise voltage trace (c) is input to it.
90 UNDERSTANDING RADAR SYSTEMS
The correlation receiver has an impulse response h (t) = u (T - t), where the
constant T simply ensures causality. In the frequency domain, this gives rise
to a transfer function
= f:", u(s)e-j,.,(T-S)ds
=e-jroTU*(w) (4.64 )
where we have assumed that u(t) is real t . Since e- jwT is simply a delay term,
we see that the essential part of the transfer function is the complex conjugate
of the Fourier transform of the transmitted pulse. This receiver transfer
function is known as the matched filter.
The response of the receiver to the signal term is given by
v(t) = u*h(t - 'rd ) (4.65 )
which in the frequency domain becomes
V(w) = e-i"'Td U(w)H(w) = e-jw(Td + 7')1 U(w)12 (4.66 )
t For simplicity, Sees 4.7 and 4.8 have assumed a real transmitted pulse u(t). When u(t) is
phase-modulated, it is normally more convenient to treat it as a complex signal. In this case,
the impulse response function giving maximum SNR is
hit) = u*(T- t)
The form of the frequency domain filter (Eq. (4.64» is unchanged.
RADAR DETECTION THEORY 91
This normally implies a loss of information about the shape of u (t), since
all the phase relations between the various frequencies have been lost. The
signal term in the output from the matched filter has a Fourier transform
that is the energy spectrum of the transmitted pulse, multiplied by a linear
phase term corresponding to range and filter delay.
Essentially, there are four steps in the treatment of signal detection given
above:
We have carried out this sequence of steps for the simplified case where
the incoming signals are at baseband. In practice, they will be at IF. This
more realistic case is dealt with in Sec. 4.11.
and Pd = 0.8. Hence the SNR on a single pulse is 1.95. This gives a
detection probability of only 0.003 for a single pulse.
For the radar to work on a higher single-pulse detection rate would
be a waste of power, given that coherent integration is possible. Multiple
pulses are an essential part in designing radar systems to meet specified
detection probabilities, while operating at reasonable powers.
(a) 0.5
,-,
I \
I \
I \
\
\
\
\
\
\
\
\
\
\
\
\
\ ,
o
(b) O.5,j8
O.4,j8
,I,,
1\
, I
, I
O.3,j8 I I
I I
I I
I I
O.2j8 I I
I I
I I
I
O.l,j8 I
I
I
I
I
,
I
\
\
2
Figure 4.12 The PDFs of noise and signal plus noise for (a) a single pulse and (b) the coherent
average of eight pulses.
As we have already indicated, a more realistic model for the radar system
and receiver takes account of the fact that radars use a modulated carrier
wave of frequency Wo on transmission, and that the signal is narrowband-
filtered on reception. After reception, the carrier is removed by an envelope
(or square-law) detector. The target detection operation is carried out on
the ensuing baseband signal. In this case, after reception, the return from a
hard (non-distributed, non-fluctuating) target would have the form
y(t) = x(t) + m(t) (4.69)
Here x (t) = v (t) cos (wot), where v (t) is a baseband modulation term
(corresponding to the baseband signal used in Sec. 4.4 et seq.), and m(t) is
narrowband noise. Such noise can be represented as
m(t) = r(t) cos [wot - tJ>(t)] (4.70)
where r(t) and tJ>(t) represent the fluctuating amplitude and phase of the
noise, both of which have bandwidths comparable to the bandwidth of the
receiver and small compared with Wo. Equation (4.70) can be expanded as
m ( t) = X ( t ) cos (wot) + Y ( t ) sin (wot ) (4.71 )
whereX(t) = r(t) cos tJ>(t) and Y(t) = r(t) sin tJ>(t) are uncorrelated gaussian
zero-mean random coefficients, with the properties
E[X2(t)] = E[y2(t)] = E[m 2(t)] = u! (4.72)
(see e.g. Schwartz2 ). Then
y(t) = [v(t) + X(t)] cos(wot) + Y(t) sin(wot) (4.73)
The detection operation uses the envelope of this signal,
r(t) = J{[v(t) + X(t)J2 + y2(t)} (4.74)
The distribution of this envelope was first derived by S.O. Rice of Bell
Telephone Laboratories 7 • It has since been discussed by many authors (e.g.
Schwartz2, Papoulis8), and has the form
where we have written V for the instantaneous value of the slowly varying
signal v(t). In Eq. (4.75), lo(z) is the zero-order modified Bessel function
of the first kind, given by
(4.77)
Pd = OO
p,(r) dr = foo exp (_V2) e- 10 (V- J2u )du
-2-
u
(4.79)
f
tam k 2 [2 2um Um
The last expression was obtained from Eq. (4.75) by the substitution
u = r2 /2u~. Since SNR = V 2/2u~, this can be written
so that Pd depends only on the SNR and k. Using Eq. (4.78), it is easy to
express k in terms of Pfa. This has been used in Fig. 4.13 on page 96 to plot
Pd as a function ofSNR for a variety of values of Pfa. We see that, as the SNR
increases, so does Pd , which indicates that matched filtering is required to
maximize Pd.
In Levanon 6 , a convenient approximation is quoted that effectively
eliminates k from Eqs (4.78) and (4.80). This yields the following direct and
easy-to-use relation between Pd , Pfa and SNR
SNR = A + 0.12AB + 1.7B (4.81 )
where
A = In (0.62)
Pea.
B=ln(~)
1- P d
and SNR in Eq. (4.81) is not in dB. This relation is fairly accurate if
10- 7 < Pfa < 10- 3 and 0.1 < Pd < 0.9
1.0
O.S
0.6
0.4
0.2
Figure 4.13 Plots of Pd against SNR for a range of values of Pra' after envelope detection.
and
(4.83 )
RADAR DETECTION THEORY 97
These are both forms of the chi-squared family of distributions. The first is
also known as the exponential PDF, and applies to a target comprising many
independent scattering centres of approximately equal strength. This distri-
bution has a standard deviation (Jav that is equal to the mean. The second
PDF is appropriate to targets that have a dominant scatterer together with
a number of other smaller scattering centres. In this case, the standard
J2.
deviation is (Jav/
The other type of variation is in the rate of change of the RCS. Again,
Swerling suggested two idealizations. The first of these is when the target
RCS changes on a timescale comparable to or faster than the PRF of the
radar, so that different pulses provide independent samples of the fluctuating
RCS. The second is when the target RCS is effectively unchanging between
pulses, but fluctuates on the timescale of the radar's scanning pattern. These
different timescales affect the probability of detection when several pulses
are combined (see Sec. 4.10). Combining the two types of PDF with the two
rates of variation yields the four Swerling models. These are normally referred
to as Sweriing cases 1-4, as set out in Table 4.2. We include the non-
fluctuating case as case 5.
Other types of fluctuation have been considered (see, e.g. Berkowitz l l ),
but the original Swerling cases still form the basis of most target detection
studies.
In order to find out how Swerling fluctuations affect the detection
probability Pd , we must follow the steps set out in Sec. 4.9. The basic problem
is the derivation of the PDF of signal plus noise, and its subsequent
integration. Swerling 10 carried out the necessary calculations for square-law
detection. In this case, if V= v(t) is the signal voltage, the variable V 2 /2
(which measures signal power) will have a PDF of the same form as the
Swerling cases, with a mean value that we will write as V~. For single pulses,
there is no difference between pulse-to-pulse and scan-to-scan variation.
Using a normalized variable
y2(t)
Z=--
2(J;'
4.82 Slow I
4.82 Fast 2
4.83 Slow 3
4.83 Fast 4
Delta function None 5
98 UNDERSTANDING RADAR SYSTEMS
pz(z)
1 (-z)
= 1 + 2So exp 1 + 2So (4.84 )
1.0
0.8
0.6
/-H---P la = 10-'
H - - - - P la = 10-<>
# - - - - P la = 10- 7
0.4
0.2
SNR
Figure 4.14 Plots of Pd for single pulses after square-law detection, as a function of SNR, for
Swerling cases 1 and 2 and a range of values of Pr••
RADAR DETECTION THEORY 99
Pd = C sJ (
+\ / 1 + 1 : So + ;J C~ ~J
exp
(4.87)
4.13 SUMMARY
Key equations
• For mean-zero gaussian noise:
p{n(t) > ka} = t- <I>(k)
• Output noise power from receiver with integrated gain Eh :
a; = NEh/2
• Noise power:
N = kTo = -204 dB W /unit bandwidth
100 UNDERSTANDING RADAR SYSTEMS
• The matched filter has a system transfer function that is the complex
conjugate of the Fourier transfer of the pulse (ignoring the delay for
causality):
H(w) = U*(w)
• Swerling PDFs:
p,,«(J) = _1 ex p ( _~)
(Jay (Jay
-1 (0.62)
A- n -
Pfa
B=ln(~)
1- P d
4.14 REFERENCES
4.15 PROBLEMS
4.1 A synthetic aperture radar image consisting mainly of grassland is of dimensions 512 x 512
pixels. The image is being scanned to find stationary vehicles thought to have an Res 10 dB
above the mean clutter level. Assuming exponential speckle, find the probability of a false alarm;
how many false alarms would be likely to occur in the image? To stress the great difference
between exponential and gaussian probabilities, find the signal-to-c1utter ratio needed to give
the same false-alarm rate if the background could be considered gaussian. (The signal-to-c1utter
ratio is the ratio of target brightness above the mean and the clutter standard deviation).
102 UNDERSTANDING RADAR SYSTEMS
4.2 (a) Show that the signal output from a correlation receiver matched to a rectangular pulse
is a triangular pulse.
(b) Two possible waveforms to be used for detection purposes are shown as (i) and (ii); (ii)
is a binary pulse generated by phase modulation. Each has its own matched filter. Are their
detection performances different? Find the outputs from their respective matched filters, and
comment on any possible consequences of the differences between them.
(i) (ii)
v
v
1 ms
-I ms I ms -1 ms
-v
4.3 What would be the expected thermal noise power at the output of a receiver whose impulse
response function is
2
COS (ttt/T) ifltl";; T/2
h(t) ={
o otherwise
when T = 10- 3 s? This receiver is matched to pulses of the same shape (because it is
symmetrical). Find the SNR at the output of the receiver if the returned signal on input is
10- 8 cos 2 (tt/T)(t - 10- 3 )
Sketch the form of the signal part of the output.
4.4 Confirm the statement in the worked example following Eq. (4.34) that a receiver with a
triangular system transfer function generates only two-thirds of the noise power of a receiver
with a rectangular passband of the same bandwidth.
How would this result change if we defined the bandwidth for the triangular system transfer
function as the bandwidth between the two points at which the system transfer function falls
to zero?
4.5 A ground-based surveillance radar with effective aperture 10 m 2 operates at a wavelength
of 25 cm, using 1 ps pulses and a peak power of 0.1 MW. Assuming a matched filter and a noise
figure of 10 dB, what would be the single-pulse detection probability from an aircraft with RCS
5 m 2 at a range of (a) 50 km and (b) 100 km, if the false-alarm rate was set at 1O- 6 ?
4.6 The signal-to-noise ratio defined in Chapter I involves bandwidth, but this does not appear
in the definition of Sec. 4.5. Explain why there is no inconsistency here.
4.7 What SNR would be needed for a 90 per cent single-pulse detection rate and a false-alarm
rate of 10- 6 using (a) the gaussian treatment and (b) the equations for an envelope-<ietected
signal?
RADAR DETECTION THEORY t03
4.8 If 16 pulses can be integrated coherently. what single-pulse detection rate is allowable in order to
pennit 80 per cent detection rate with a false-alann probability of 1(}-'5?
4.9 We will meet the chirped pulse
exp( -jat2) ifltl..: T
u(t) ={
o otherwise
(where a is a constant) in Chapter 6. Find its matched filter. (You will need the footnote
following Eq. (4.64).) •
4.10 For 80 per cent detection rate and 10- 6 false-alarm rate, what peak power would be
acceptable in a radar coherently averaging thirty-two 2 JIS pulses, if it is operating on a 24 ern
wavelength with an effective aperture of 15 ml, has a loss factor of 50 per cent and needs to
locate aircraft with RCS down to 20 m l within a range of 100 km?
CHAPTER
FIVE
SIGNAL AND DATA PROCESSING
Separating signals from clutter, and tracking targets have become sophisticated
arts.
5.1 INTRODUCTION
The reliability and low cost of modern digital electronics have revolutionized
radar engineering. Over the years, the AID converter has moved progressively
up the receiver chain towards the antenna, increasingly replacing analogue
electronics by digital hardware and computer software. This engineering
activity has two objectives: to remove unwanted clutter as far as possible,
and to detect targets against the residual clutter and noise.
Radar engineers make a distinction between signal processing, the fast
hardware processing developed mostly by electronic engineers, and data
processing, which is concerned mainly with target-detection software and is
very much the preserve of mathematicians. The dividing line between the
two is neither clear-cut nor stationary, as it gradually moves up the receiver
chain with the increasing ability of modern computers to replace hard-wired
circuits. The current position for a typical radar system is shown in Fig. 5.1.
The objective of this chapter is to take a guided tour through the signal
processing and data processing shown in the block diagram of Fig. 5.1, but
104
~ r---,
r-l MTI f-..,
I L _ _ ..J I Track initiation
I ,--------, I
Plot extraction
Plot-track
association
Signal processing
Data processing
~
106 UNDERSTANDING RADAR SYSTEMS
we cannot begin until we know more about the properties of the clutter we
are trying to filter out. The simplistic case might be the problem of detecting
an aircraft flying against a bare mountainside in the background. In this
case the clutter is stationary. But suppose the mountainside is covered in
trees blowing about in the wind, and the aircraft is a helicopter nearly
hovering; our problem becomes more difficult. Perhaps the ultimate challenge
is presented by an air-borne radar in look-down mode trying to distinguish
a tank from wind-blown vegetation and moving clouds of rain. Clearly it
becomes important to study the properties of ground clutter before we begin
signal processing.
Pe(c)
C
= 2"exp
Ue
(_c
--2
2ue
2
)
(5.1 )
The expressions for detection and false-alarm probabilities are exactly the
same as those derived in Sec. 4.11, with clutter taking the place of noise.
This clutter model is valid as long as, within the illuminated region, the
clutter can be regarded as consisting of many independent random scatterers
of comparable strength. Such conditions normally hold in weather clutter
or chaff, where the scattering is from very many droplets or particles. They
are also encountered in other common situations, such as low-resolution
marine radars operating in calm sea conditions, or radars illuminating desert
or agricultural areas. When the radar is operating under conditions where
the return may be dominated by a few brighter scatterers within each
resolution cell (such as can happen when a high-resolution radar illuminates
a patch of ocean whose dimensions are comparable with the water wave-
length), then this model no longer holds. Under such conditions, larger
values of clutter occur more often than expected from the Rayleigh distri-
bution, so that false alarms will occur more frequently unless the detection
threshold is raised.
Analysis of the detection/false-alarm rates for non-gaussian clutter
requires a model for the clutter and signal-plus-clutter PDFs at the detection
stage. In practice, this normally relies on empirical distributions, supple-
mented by physical reasoning. Several types of model are in use for different
circumstances. For sea clutter, log-normal, Weibull and K distributions have
been used inter alia. With different parameters, the same models also have
SIGNAL AND DATA PROCESSING 107
their place in describing land clutter from a variety of terrain types. (See
reference 1 for a description of these distributions and for further references. )
When the targets being sought by the radar are moving, doppler
information can also be used to separate them from the clutter. This is
complicated by the fact that many forms of clutter, such as the sea surface,
wind-blown vegetation, chaff and rain clouds, are themselves in motion, and
hence also give rise to doppler shifts. As a result, the clutter spectrum is
normally spread about zero doppler frequency. For most purposes, the
spectrum C (w ) can be treated as having a gaussian shape, i.e.
Most radar systems need some form of doppler processing to filter out clutter
and thereby reveal faster-moving targets. These days, such filters are
implemented digitally, either as some form of fast Fourier transform (FFT)
algorithm or as a set of transversal filters.
108 UNDERSTANDING RADAR SYSTEMS
Here the a, are the weighting constants. Digital approximations to, for
example, the matched filter described in Chapter 4 can be constructed in
this way. (There are many good books dealing with this topic; Stremler 2 is
an example.)
A digital approximation to the Fourier transform can also be constructed
in a similar way, using complex weights. This is the discrete Fourier transform
(DFT).1t is related to the sequence {I,}~(/ by the equation
M-l
Fp = L I, e - 21!jlp/M for 0:::::; p:::::; M - 1 (5.4 )
1=0
Delay = I pulse
interval T
$
110 UNDERSTANDING RADAR SYSTEMS
,---Simple canceller
~---->'...--- Improved canceller
~-------------=-"';::"'-Doppler frequency
Slow targets Fast targets
would be complete in the absence of noise. If the echo has changed phase
slightly (and amplitude strictly, but limiting amplifiers can be used to remove
the amplitude dependence) because of its motion, then cancellation will be
less complete. For a target in uniform motion there is a constant change in
phase from pulse to pulse and cancellation does not occur. The two-pulse
cancelling MTI is therefore acting as a high-pass filter, as shown in Fig. 5.3.
The impact of two-pulse cancellation on the clutter can be quantified
by the clutter attenuation factor
CA = input clutter power (5.5 )
output clutter power
We saw in Chapter 4 that noise power and noise variance are the same thing,
and that we can calculate noise power by integrating the power spectrum
(see Eq. (4.23». We also saw how to calculate the output spectrum given
the input spectrum and the system transfer function (see Eq. (4.28 Exactly ».
the same principles can be applied to clutter. This gives
CA = --:------=-J_C_(£0_)d_£O---=--_ (5.6 )
Jc(£O)IH(£OW d£O
where H (£0) is the system transfer function of the canceller. Since the output
of the canceller from an input f (t) is given by
g(t) = f(t) - f(t - T) = f(t)*[b(t) - b(t - T)] (5.7)
where Tis the pulse separation, its impulse response function is b(t) - b(t - T),
and hence
H(£O) = 1 - e- jwT = 2j e- jwT/ 2 sin(£OTj2) (5.8)
Using Eqs (5.2) and (5.8) in Eq. (5.6) gives, after simplification,
CA = _ _ _ _0_.5_ _ __ (5.9)
SIGNAL AND DATA PROCESSING III
(a)
(b)
From
PSD
+-- Multiply by
coefficients
Figure 5.4 (a) A three-pulse cancelling system. (b) A more general multipulse cancelling
scheme.
Transversal filters and MTIs are often modified to null out moving clutter,
such as rain clouds, with a clutter notch that can be controlled adaptively.
The whole of MTI thinking is geared towards removing clutter in this way,
so that a target is left competing only with system noise.
A different approach is to concentrate on target detection by developing
a special filter that allows only the target through. The target would then
be detected in the narrowest possible bandwidth, so minimizing the noise.
Unfortunately, until the target has been detected, we do not know its speed
and we cannot construct the filter. The solution to this dilemma is to use a
bank of filters, just overlapping, such that the target must appear in one of
them. The slow-speed filters will be contaminated with clutter, but we can
search through the remainder to find the target. The output of this filter bank
is known as the doppler spectrum (see Fig. 5.5).
There are many ways of constructing a filter bank, including analogue
methods, but these days the quickest way is to use a FFT algorithm. To
do this, a set of samples {u/} from a fixed range gate are gathered, using the
multiple pulses available within a single scan. The effect of the FFT is most
easily explained if we assume a non-fluctuating target with doppler frequency
Wd' We can therefore write the signal sequence as
(5.11 )
SIGNAL AND DATA PROCESSING 113
Frequency [Hz)
Figure 5.5 The output of the FFT can be thought of as a bank of bandpass filters.
where T is the time spacing of the samples. The amplitude of each sample
is the constant V, and for simplicity we have ignored the carrier frequency.
The DFT of this sequence (using Eq. (5.4)) is
M-l M-l
Up =V L exp(jwd/T) exp( -2nj/p/ M) = V L exp(j(wdT - 2np/ M )/]
1=0 1=0
(5.12)
Writing
z = exp[j(wdT- 2np/M)] (5.13 )
we can see that this is a geometric progression. Hence we can write down
the sum as
1- ZM
U =V-- (5.14 )
p 1- z
Since ZM = exp(jwdMT), the quantity on the top line of Eq. (5.14) is
constant; it does not depend on p. Hence as p varies, Up is maximized when
we make the bottom line of Eq. (5.14) as small as possible, i.e. set z = 1.
Using Eq. (5.13), this means that the maximum term in the FFToccurs when
(5.15)
From this, we can find the doppler frequency Wd'
Equation (5.15) implies that the possible solutions for the doppler
frequency are also digital (they are all multiples of 2n / MT). Hence there
may be some error introduced in the estimate of the doppler frequency, with
the correct value lying between the maximum and one of its neighbours. At
this point, you may object that this error is needed to stop the expression
in Eq. (5.14) becoming infinite when z is exactly 1. This does not happen,
because if z is exactly 1, the summation in Eq. (5.12) gives the value MV.
(You should check this.) So the maximum value in the FFT is MV. (It is
also worth noting, if the argument above makes you uneasy, that the limiting
value of Eq. (5.14) as p gets close to wdMT /2n is MV.)
114 UNDERSTANDING RADAR SYSTEMS
To find the SNR improvement from this process, we can write the input as
I, = u, + n,
where u, and n, are the signal and noise terms in the input sequence. The
signal terms have the same form as above, and we assume that the noise
terms are all mean-zero, uncorrelated and with variance <1;. Taking the DFT
gives
M-l M-l
Fp= L u,exp(-2njlp/M)+ L n,exp(-2njlp/M) (5.16 )
'=0 '=0
The first summation is deterministic, and is given by Eq. (5.14), but the noise
summation is random. We need the average noise power, where the noise
term is given by the sum in Eq. (5.16), i.e. by
M-l
Np = L n,exp( -2njlp/M) (5.17 )
'=0
This requires a slight modification of the methods used in Chapter 4, since
the noise term here is complex. In this case, the noise pow,er is found by
averaging the squared modulus of N p , so that
Noise power = E[INpI2] = E[NpN;] (5.18)
Therefore
M-l M-l ]
Noise power = E [ ,~o n, exp (2njlp / M) k~o nk exp ( - 2njkp / M)
M-l ]
= E [ Ii.to n,nk exp[ -2njp(l- k)/ M]
M-l
= L E[n,nk] exp[ -2njp(l- k)/ M] (5.19)
k.'=O
signal energy for each sample is V 2 , and the SNR is therefore given by
SNR (before) = V2/20';
At the maximum response of the OFT, however, the signal gives a value
of M V. Hence the SNR is
SNR(after) = M 2V 2/2MO'; = MV 2 /20';
Thus there is a gain in SNR by a factor M at the peak response of the 0 FT.
The analysis has only considered the case of uncorrelated noise. For
clutter, we would need to take account of the correlation of the clutter
samples (including the phase correlation, which gives rise to the doppler
spectrum of the clutter). This complicates the analysis, and is not carried
out here. However, in the range of doppler frequencies where the clutter
contribution is negligible and noise is the dominant factor, the analysis carried
out above is valid.
5.5 THRESHOLDING
Signal
~
Q:l
~ Noise
t I-----~...------_f__+_---------_r- Threshold
~
~
- Mean
L -_ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ ~~--~~__+
Time/range/doppler
be 7.2 dB. This addition of a margin in dB implies that, as the mean noise
level varies, the threshold also varies in such a way as to keep a constant
false-alarm rate, as long as the noise model described in Sec. 4.11 is valid.
In the case of some rarer noise models, a more sensitive detection system
can be designed (for the same false-alarm rate) by using other ways of creating
a threshold, such as the linear addition of a margin to the mean noise power.
There are only two main decisions to be made in thresholding: the first
is the choice of the basis on which to calculate the mean power, and the
second is the setting of the margin to achieve the correct sensitivity.
The mean power is calculated by finding the average power in reference
cells, which are near the test cell, but which are not so close that they are
contaminated by the presence of the target. The usual procedure is to leave
guard cells either side of the test cell, which are not included in the mean,
as shown in Fig. 5.7. The number of cells contributing to the mean, and the
width of the guard band, depend on how quickly the background is changing
and how much the target is spread across the cells. The method must therefore
be optimized for each radar system, but typical schemes are shown in Fig. 5.8.
So far we have not explained what we mean by a cell adjacent to the
test cell. It could be a radar resolution cell close to the test cell in range,
doppler, azimuth, elevation or even time (i.e. what happened in that
particular cell in previous scans). All of these parameters have been used at
one time or another, but one-dimensional thresholding in the range dimension,
or two-dimensional thresholding in the doppler/range plane are probably
the most commonly used schemes.
Having found a method of determining the mean, the next task is to
add a suitable margin. This is the problem we analysed in Chapter 4, where
we derived the relationship between SNR, the margin k and the false-alarm
and detection probabilities. We showed that, if the margin is set very low,
SIGNAL AND DATA PROCESSING 117
r---L----, Decision
Figure 5.7 Guard cells G are used to prevent large signals in the test cell T from contaminating
the mean of the reference cells R.
Time
---+
Ranger
RR RR R RR RR R
RR R T RR R
RR RR R RR RR R
---+
Doppler
Figure 5.8 Various schemes for finding local reference cells to determine the mean level around
the test cell.
the probability of detection will be good (even faint target echoes are likely
to be detected) but the probability of false alarm will also be high because
a significant number of noise spikes will trigger the system (see Fig. 4.8).
Frequent false alarms can be very inconvenient if they trigger a defence
118 UNDERSTANDING RADAR SYSTEMS
system, but there is an even greater danger. Too many false alarms will
swamp the tracker and cause the whole data processing computer to crash;
this could temporarily paralyse a defence system completely. As a result, low
false-alarm rates are used; the value for PCa of 10- 6 that we have adopted
throughout this book is comparatively large for many systems.
The solution to the problem of setting the margin correctly is to leave
the job to the computer. If there are too few threshold crossings, the margin
can be lowered to make the system more sensitive, so that the system is
triggered more frequently. If the number of threshold crossings increases too
far, exceeding a set limit, the margin is automatically increased to control
it. To the beginner this seems a rather empirical approach to the problem,
which leads to the probability of detection varying with the level of
clutter/noise/jamming. In practice, there really is no other way to proceed,
except possibly in the most advanced computer systems where more
processing power can be allocated to those areas where higher clutter levels
are experienced. Keeping the number of threshold crossings constant is
effectively the same as keeping the false-alarm rate constant (assuming that
very few of the crossings are due to real targets). Hence this technique has
become known as constant false-alarm rate (CF AR) processing and is widely
used in radar practice. If the false-alarm rate is kept constant and the clutter
statistics change, it is inevitable that this leads to changes in the probability
of detection. This is because changes in the clutter power are equivalent to
changes in SNR; Eqs (4.45) and (4.80) describe quantitatively how this
affects Pd.
The CF AR process can be improved to cope with contamination
of reference cells by clutter or other targets. Sometimes it is considered
worthwhile to make a second pass through the data after the largest signals
have been identified and removed by the first pass. This double thresholding
improves the sensitivity of the system, but at the cost of increased pro-
cessing effort.
Once a potential target, or plot, has been identified by a signal crossing the
threshold, the next step is to extract all the available information about it.
Azimuth and elevation information are usually extracted by monopulse
techniques, and the range by interpolation between range gates, as described
in Chapter 3. Similarly, the target speed is found by interpolation between
doppler cells.
The purpose of collecting this information at this stage is because a true
target will behave in a fairly predictable manner in each of the range/angle/
speed dimensions and will be said to 'track' correctly. Plots arising from
noise or clutter spikes (false alarms) usually behave randomly and do not
SIGNAL AND DATA PROCESSING 119
track; for example, the rate of change in range will not agree with the velocity
estimate obtained from the doppler information. However, this is not always
the case; on occasion, sea clutter spikes due to persistent waves may recur
sufficiently often for false tracks to be initiated.
Existing track 1
Association gates
~
~
Track 2 Track 2
Track 2
Figure S.10 There are potential complications when tracks appear to cross.
plot A go with track 1 and B with track 2? The answer depends on our
track model. If the two targets were ships, we would assume this association
to be correct, as they might be expected to diverge, as shown in Fig. 5.lOb;
but if the targets were aircraft at different heights, our belief might be that
they continued as straight tracks, as in Fig. 5.1Oc.
Other information is also contained within the track model. For example,
the plots should be distributed randomly either side of a straight track,
assuming the observations are corrupted by zero-mean gaussian noise. The
run sequence of '1 's (above the track) and 'O's (below) can be tested for
randomness and used to help decide which plot best fits with which track.
Simple track initiation schemes might use + 3 for a hit, - 2 for a miss
and declare the target to be present when the count exceeds 7. A weak, fading
target might therefore take some time to be detected as it approaches the
radar, as the following sequence shows:
1 H +3 3
2 M -2 1
3 H +3 4
4 M -2 2
5 H +3 5
6 M -2 3
7 M -2 1
8 H +3 4
9 H +3 7
10 H +3 10 Track declared present
Some of the limitations of such simple schemes can be avoided by the Boolean
combination of rules such as 'seven out of ten scans must be hits AND two
of the first three must be hits'.
Track initiation schemes tend to be empirical and they turn out to be
quite hard to compare analytically, because of the Bayesian nature of the
problem. In the long term they may be replaced by artifical intelligence
methods of searching for patterns in the data (see Chapter 15).
How should you decide whether a plot is near enough to a track to be
included as a possible candidate for that track? The answer is shown in
Fig. 5.11. The first plot defines the position of the target, a civil airliner for
example, at a moment in time. Ten seconds later the radar returns to scan
the same sector, and the aircraft could be anywhere within a circle of radius
equal to the maximum possible speed of the aircraft times the 10 s interval,
giving a typical radius of 3 km. After the second plot has been identified
Figure 5.11 The plot-track association gate decreases with increased observations.
122 UNDERSTANDING RADAR SYSTEMS
within this circle, there is some indication of the aircraft heading, so the next
association gate need not be so large. As Fig. 5.11 shows, the association
gates can eventually be reduced to a size determined by the accuracy of the
tracker. Usually, this gate size is about one standard deviation for a straight
track, but nearer 3 standard deviations for a manoeuvring target.
5.9 TRACKING
Measurement
~.
• • • ---_--~ Predicted position xp
Smoothed track Xn xp •
• •
• • Measured position x"
a = 1 ~ Joins measurements
number between 0 and 1. When a = 1 we are joining the dots, and when (X = 0
we are ignoring the data. In this way we can use (X to define the new smoothed
position of the target Xs to be
(5.24)
(5.25 )
f3 = 6 (5.28)
n(n + 1)
where n = number of observations.
Putting a few values of n into Eqs (5.27) and (5.28) shows that there is
almost no smoothing of the initial data, but the damping factor is pro-
gressively increased as the observations continue. This increased confidence
in the true track of the target means that the tracking error is reduced with
increasing time.
The error variance of the smoothed positions, which tells us the tracking
error, is given by
2 2(2n - 1) 2
(1 = (1 (5.29)
n n(n + 1)
The error variance of the predicted position, used to set the size of the
association gate, is given by
2(2n - 1)
(1;I(n- J) = (12 (5.30 )
(n-1)(n-2)
Two independent trackers are used in the x and y coordinates and the
algorithm can also be extended to track in acceleration in an (X - f3 - y
124 UNDERSTANDING RADAR SYSTEMS
Actual
target track
Measurements
/~
...L.._..,...._.%...."""T'"_...L.._~_"T'"""_~:........:L---L---L...J.._ _ Forecast track
uncertainty (and hence error variance) in the track model when the larget
is manoeuvring.
Another problem is to decide what to do if the track appears to branch.
Track branching occurs when there are so many false plots that the true one
cannot be identified with certainty. Some systems attempt to cope with this
problem by tracking both branches for a while and then trying to select the
most promising looking branch; this is not unlike the Viterbi algorithm used
in digital communications to find the decoding route through the trellises
used in convolution encoding. Another approach is to use probabilistic data
association (PDA), formulated originally by Bar-Shalom and Tse 4 and
extended by others to take account of multiple targets, manoeuvres, track
initiation, etc. (see e.g. Colgrave et al. 5 ). PDA builds on the Kalman- Bucy
filter to use an average of all the validated plots, weighted by the probability
of their having originated from the target. Although PDA introduces
inaccuracies from the (weighted) inclusion of incorrect plots, it avoids the
risks of an increasing number of track branches eventually jamming the
computer and causing a system crash.
If the target track enters a partic~larly dense area of clutter or jamming,
it may not be possible for the software to maintain the track and operator
intervention may be necessary. Part of the effectiveness of radar systems is
determined by the skill, intelligence and experience of operators. In future,
it may be possible to use expert systems to replicate the local knowledge of
experienced operators, but it is unlikely that high-level decisions in radar
(e.g. whether to fire on a target) will be taken without .human control, or at
least a human right of intervention.
5.10 SUMMARY
The target detections that initiate a track must take advantage of whatever
gains are possible in SNR, because viable radar systems must operate at
126 UNDERSTANDING RADAR SYSTEMS
Key equations
CA _ ( 2c )2
woTvrms
2(2n - 1)
a.=----
n(n + 1)
fJ = 6
n(n + 1)
where n = number of observations.
5.11 REFERENCES
5. Track initiation and nearest neighbour incorporated into probabilistic data association,
S. B. Colgrave, A. W. Davis and J. K. Ayliffe, IE Aust. IREE Aust., 6(3),191-198, 1986.
5.12 PROBLEMS
5.1 If the signal and noise were roughly equal at the output of a radar receiver, what SNR
would you expect after averaging four 32-point FFT blocks of observations? Assume the target
RCS and speed to be constant. What would happen to the SNR if the target began to manoeuvre?
5.2 A surface-wave HF radar uses a PRF of 275 Hz and observes an aircraft for 3.72 s. What
is the potential coherent gain and how would this improve the range estimate? These radars
use vertical polarization (see Chapter 10). Would the target remain visible through a manoeuvre?
5.3 A radar system is required to have a high probability of detection of95 per cent on envelope
detection of a single pulse. This Pd is to be achieved at the expense of a high false-alarm rate of
10- 5 • What margin should be chosen for the thresholding and what.SNR is required?
5.4 The following series of x-coordinate position measurements of a target were made at the
rate of one per second:
1 2 3 4 5
x. o 35 88 118 158
Tabulate the values of IX, p, Xs and v. for the IX-P tracker and predict the x value for the next
position.
5.5 In problem 5.4, what improvement in the target position results from the track smoothing?
Suggest a size for the association gate in the x dimension for the sixth ·observation.
CHAPTER
SIX
DESIGNING RADAR WAVEFORMS
6.1 INTRODUCTION
t Throughout we assume no bias in the measurement, so that accuracy and precision will
be equivalent.
128
DESIGNING RADAR WAVEFORMS 129
Ambiguity occurs if the output of the receiver from a single target contains
multiple peaks that can be mistaken for other targets. Such peaks may be
caused by noise, but may also be produced by the shape of the transmitted
waveform. An obvious example is if there are significant sidelobes in the
radar antenna pattern. Objects illuminated by the sidelobes will be interpreted
as though they were in the main lobe. This will lead to angular positioning
errors or mUltiple detections generated by a single target. It can also cause
clutter to be interpreted as target.
Accuracy refers to the expected spread of measurements about the true
value. We have seen that detection and ranging are essentially the same
problem. The range of a target depends on the time delay at a peak in the
output from the matched filter, as long as this peak exceeds some threshold.
Noise in the output will cause the peak to be displaced randomly from its
true position. The standard deviation of this variation will be our adopted
measure of range accuracy. We adopt a similar definition for the accuracy
of measurements of doppler frequency.
All these aspects of system behaviour are affected by the radar bandwidth,
whose definition we need to make more precise. Several different definitions
of this important concept are in use. The simplest definition is applicable if
the signal u (t) is band-limited, i.e. it has a Fourier transform U (w) for which
U (w) = 0 if w > Q rad s - 1. Then it is natural to take the bandwidth as
Q rad s - 1, or Qj2n Hz. None of the commonly used pulse modulations satisfy
this relation exactly, but it is of considerable theoretical value because of its
relation to the Shannon-Whittaker sampling theorem t (band-limited signals
can be reconstructed exactly from their samples as long as the sampling
frequency exceeds the Nyquist rate of Qjn samples per second.) In practice,
many real signals are deliberately band-limited by the use of anti-aliasing
filters.
The most commonly used engineering definition of bandwidth is the 3 dB
width, which is the separation in frequency of the half-power points in the
energy spectrum of u (t), i.e. it is obtained by solving
IU(wW = !IU(OW (6.1 )
t This chapter uses a number of results from Fourier transform theory. A good reference is
Bracewell l .
130 UNDERSTANDING RADAR SYSTEMS
J
This definition assumes that rolU( ro )l2dro == 0, which can always be
arranged. (For real signals, it will always be the case, since then IU (w)1 is
an even function of w; more generally, since u (t) is in fact the modulation
of a carrier frequency, it will arise by defining a suitable centre frequency
for the modulated signal.) This means that a;
is the second moment of the
J
Fourier transform of the unit energy signal u (t )/ E about its mean, 0. It
is therefore a measure of the spread of energy in the spectrum of the signal,
and is analogous to the variance in probability theory or moment of inertia
in mechanics. For calculation purposes, it can be very convenient to use the
Fourier transform pair
u' (t) = du/dt +-+ jwU (w)
Then using Parseval's theorem, we can write
f~oo IU'(wW dw
(6.4 )
21tE
In Eq. (6.4) we have assumed that the centroid of the pulse is at 0, i.e.
Jtlu(tW dt = 0, and we have used the Fourier transform pair
tu(t)+-+jU'(w)
Though the definition of effective bandwidth and pulse duration may
seem unnecessarily complicated, they are of considerable value because they
can be handled analytically. From them, we can derive a number of
fundamental relations between accuracy, resolution and ambiguity that
clarify our understanding of the inherent limitations of any radar system,
DESIGNING RADAR WAVEFORMS 131
We saw in Chapter 4 that the output of the matched filter consists of two
terms, one due to the correlation of the signal with itself, and a noise term
due to the correlation of the signal with the noise. In the absence of noise,
the peak of the signal term would give the true time delay corresponding to
the target's range. However, the noise term will cause this peak to move
around. In fact, if the gradient of the noise term is not zero at the correct
time delay, the peak will be displaced. As a result, the measured value of
the time delay of the peak can take a range of values. Woodward and Davies 3
c: Table 6.1 Fourier transform pairs and values of E, (1", and (1t for some basic pulseshapes
N
(a)
3 T3 + 2aT2 + 3a 2T + 4a 3
(T+a) sa(T+a)~) sa(T-a)~) t(2a + T)
(2a + T)(T - a) IO(2a + T)
(b)
2T Sa(Tw) 2T 00 T2/3
(c)
T
Cd)
cos 2 (Trtl2 n
T 3T n2 2T2 (n2/2 _ 3)
1 _ (Tw/n)2 Sa(Tw) 4 3T 2 3n 2
(e)
exp (-t 2I2T 2 )
~
-
134 UNDERSTANDING RADAR SYSTEMS
succeeded in showing that for large SNR the PDF of the measured delay
would be approximately gaussian with a mean value occurring at the true
delay and with standard deviation
1
J'td=
21turo
J (2E/ N) [s] (6.5)
6.4 RESOLUTION
t These two equations have already been encountered as the range error and doppler
frequency error (in Hz) of Eqs (1.19) and (1.22).
DESIGNING RADAR WAVEFORMS 135
1. E = Ip,,(O)1 ~ Ip,,(t)1 for all t, i.e. the ACF is maximum at zero lag, and
its value there is the signal energy.
2. For finite energy pulses, p" (t ) will tend to 0 as t increases, though there
may be subpeaks (ambiguity peaks) in addition to the peak at O.
3. p" (t) +-+ IV (w W, i.e. the Fourier transform of the ACF is the energy
spectrum of the waveform.
These properties of the ACF have led to such definitions of time resolution
as the 3 dB width of p" (t), or the time over which the ACF drops to a value
(1 Ie )p,,(O), or the time to the first zero of the ACF. These have their uses
for particular types of ACF. A measure of resolution that is more general
and is better for analysis is
A, = JIp,,(tW dt [s] (6.9)
d p;(O)
136 UNDERSTANDING RADAR SYSTEMS
1.0
0.8 TO = 0
T[ = 0
0 1.5 2.0
1.0
0.8 TO =0
T[ = 0.25
0.2
0 2.0
1.0
0.8 TO =0
0.6 T[ = 0.50
1.0
0.8 TO = 0
0.6
T[ = 0.75
0 1.5 2.0
1.0
0.8 TO =0
0.6 T[ = 1.0
Figure 6.1 The output from the matched filter when two targets with range delays '0 and, \
are present, for different values of '0 - '\. Here '0 is set to 0 and, 1 is marked on each successive
plot.
DESIGNING RADAR WAVEFORMS 137
-l
I
I
I
I
I
I
I
I
I
I
I
I
I
0.4
-4 4
Figure 6.2 The relation of the equivalent-rectangle resolution to the area under the graph of
Ip.(tW and its value at O.
This definition has a very simple interpretation (see Fig. 6.2). It is the width
of a rectangle of height p;(O) that has the same area as is under the curve
IPu(t W. Hence it is sometimes called the equivalent-rectangle resolution.
Using properties 1 and 3 of ACFs and Parseval's theorem, Eq. (6.9) may
also be written as
[s] (6.10)
This form of the definition also has a useful and instructive interpretation.
If U (co) took only the values 1 or 0, and the total length of the intervals
in which it took the value 1 was F, then F could be considered as a measure
of occupied bandwidth, or what Woodward 2 calls the frequency span. The
energy of this signal would be F12n, and hence for this signal d'l'd = 2n /F.
This expresses the reciprocal relationship between resolution and bandwidth.
Similar definitions for the frequency resolution are possible. We first
need the frequency-domain A CF
Equations (6.9) and (6.10) apply when the targets to be resolved are known
to be stationary, while Eq. (6.12) applies when targets are at the same range.
However, if we do not have such prior knowledge of the target character-
istics then we need to worry about the combined effect of a shift in range
and frequency. This can be analysed by considering the behaviour of the
output of the correlation receiver when the input is doppler-shifted. (This
means that the filter is not properly matched to the incoming signal, since
it does not replicate its frequency behaviour.) If the transmitted signal is a
complex modulation u (t) of a carrier frequency Wo rad s - 1, and the target
velocity causes a doppler shift in frequency of Wd rad s - 1, then the correlation
operation will give as output at time t
X(t,w d ) = - 1 foo
U*(v)U(v - wd)exp(jvt)dv (6.16 )
2nE - 00
The cuts across the ambiguity function along the delay (t) and doppler (Wd)
axes are directly related to the time-domain and frequency-domain ACFs (and
hence to the time and doppler resolutions), since
x(t,O) = Pu(t)/E (6.17 )
DESIGNING RADAR WAVEFORMS 139
and
(6.18 )
~
6
J: 0
:=
-3
-6 ~
Figure 6.3 The modulus of the ambiguity function of a rectangular pulse (T = 1.0).
(a)
1.0
X 0.5
(b)
whr
-0.5
Figure 6.4 Cuts along (a) the delay and (b) the doppler axes of the ambiguity function plot
of Fig. 6.3.
DESIGNING RADAR WAVEFORMS 141
1 foo
~E2 _oolu(sWds foo
_oolu(s-t)exp(-jwdsWds
1 foo 2 foo
= E2 -00 lu(s)1 ds -00 lu(s - tW ds (6.22)
Each of the integrals on the right-hand side is equal to the energy of the
signal, so that
(6.23 )
Perhaps the most remarkable fact about the ambiguity function is that, if
we view \X(t, wd)1 2 as a surface, then the volume under this surface is always
1, irrespective of the shape of the waveform, i.e.
(6.24 )
2.5
2.0
1.5
1.0
0.5
-; 0.0
-0.5
-1.0
-1.5
-2.0
-2.5
-10 10
Figure 6.5 A contour plot of the ambiguity function of a gaussian pulse (T = 2.0).
DESIGNING RADAR WAVEFORMS 143
M
, -_______________________ A~ ______________________ ~
o 0 0 0····0
Tr
Figure 6.6 A coherent pulsetrain with PRF T,. containing M pulses of width T.
- -
IXM (t ,Wd )1 - 1 M~l ISin[T.(M -IPI)W d /2]\1 Xl (t
L...
- T.
P r,Wd
)1
M p=-(M-I) sm(T.wd/ 2 )
(6.27)
where XI (t, w d ) is the ambiguity function of a single pulse, given by
Eq. (6.19). A contour plot for the case M = 3 is given as Fig. 6.7, and has the
'bed of nails' structure characteristic of the ambiguity diagrams of pulsetrains.
The system designer can control both T and T. in order to ensure that targets
of interest only occur near the central peak of this ambiguity diagram, giving
enhanced range and doppler accuracy and effective resolution.
The cuts along the delay and doppler axes are given in Fig. 6.8. Along
the delay axis Eq. (6.27) takes the form
if It - pT.1 < T, and is zero everywhere else. The term in the second
parentheses corresponds to a triangular peak of base width 2T, with its centre
at PT.. The term in the first parentheses corresponds to a triangular weighting.
This is shown in Fig. 6.8a. The triangles are spaced at intervals T., and the
condition T. > 2T is to prevent these triangles overlapping; if they do so,
the expression in Eq. (6.28) becomes more complicated.
Along the doppler axis, Eq. (6.27) has the form
3r-----------rn~.----.~~----_rrrrTr----------_,
0 10 000 0 10
2
00 88 00
CJ CJ
(0) CgJ
C@) C@)
) 0
~ @j)
C@) C@)
-1 (Q) (Q)
C) o
-2 00 00
000 000
Figure 6.7 A contour plot of the ambiguity function for the pulsetrain shown in Fig. 6.6
(M = 3, T = 1.0, T, = 4.0).
(a)
(b)
wl-rr
Figure 6.8 Cuts along (a) the delay and (b) the doppler axes of the ambiguity function plot
of Fig. 6.7.
DESIGNING RADAR WAVEFORMS 145
The radar designer's ideal waveform would give good performance as regards
the following:
• Target detection
• Range and doppler accuracy
• Range and doppler resolution
6.8 CHIRP
The simplest form offrequency modulation is linear FM, i.e. a pulse described
by
u(t) = a(t)exp(jnkt 2 ) (6.30)
Ix(t, wd)l = (
1 _!!l)
T
sa(T-ltl)(Wd +
2
2nkt))1 ifltl ~ T
(6.32 )
1
1o otherwise
The shape of the ambiguity function is not apparent from this equation, but
contours are plotted in Fig. 6.10, and cuts along the delay and doppler axes
are shown in Fig. 6.11. The most obvious feature of the ambiguity diagram
(a)
v
~~
{\ .75-
.50r
0.25r
n
H~:25 Vi
I I J J
4 - f--
2 3 4
t
0.50 r
-0.75 t-
V 1.00 ~ V
(b)
w
-4 4
Figure 6.9 (a) The in-phase component of a linearFM signal and (b) the instantaneous
frequency of this signal as a function of time.
DESIGNING RADAR WAVEFORMS 147
16
12
) 0
-4
-8
-12
Figure 6.10 A contour plot of the ambiguity function of a linear FM pulse (T = 1.0, k = 5).
(a)
1.0
0.5
-0.5
(b)
whr
-0.5
Figure 6.11 Cuts along (a) the delay and (b) the doppler axes of the ambiguity function plot
of Fig. 6.10.
148 UNDERSTANDING RADAR SYSTEMS
for small O. The minus sign is present because, in our system of coordinates,
if 0> 0, the distance x is negative (see Fig. 6.12). Since x = Vt (taking the
scatterer at the origin of the x axis, and measuring time from the instant
when the scatterer is broadside to the platform)
2V 2
fd = - - t [Hz] (6.35)
2R
This is a linear FM signal like that described by Eq. (6.31), for which
k = - 2 V 2 /2R. The SAR processor compresses this signal, and for a
stationary scatterer the peak response will occur in the right place in the
image relative to other stationary scatterers. If the scatterer is moving and
has a velocity component along the line of sight to the radar, the associated
doppler shift causes the response from the scatterer to be a cut across the
ambiguity diagram parallel to the delay axis, but moved up or down. As we
can see from Fig. 6.10, this moves the maximum response to a later or earlier
DESIGNING RADAR WAVEFORMS 149
-----r4h, [ \
~
~ F
[
[
Flight track
•
I6\ I
I \ I
I \ I
[\ I
I \ I
[ \ I
I \ I
I \ [
[ \ :
\ [
\ I
\6 1
\i
, I
\ I
\ [
------------------~,~----------------. x=O x
where we have used the notation fdu to indicate the excess doppler frequency
uncompensated for in the SAR processing. Since the maximum response
occurs at the wrong time, the scatterer will be misplaced in the image. The
positional error corresponding to the time error of Eq. (6.36) is
RA
Vte = 2V fdU [m] (6.37)
Because of this, SAR images often show effects such as ships displaced from
their wakes, or cars apparently in the middle of fields instead of on the road
along which they are travelling. The magnitude of the displacement (which
is in the along-track direction, even though caused by cross-track motion)
can sometimes be used to estimate the velocities of moving scatterers.
SOLUTION Since fdu = 2Y.1 A, where y. is the radial velocity, the dis-
placement is given by
d=R(Y.IV) [m]
Here Y. = - 20 cos 85° ~ - 1.74 km h - 1, so that the displacement is
approximately 348 ffi.
COMMENTS It is clear that quite small doppler shifts can cause large
apparent displacements in the image. In fact, the calculation is compli-
cated by the fact that SAR is sampling the signal. Since the bandwidth
of the FM signal used in the along-track processing is typically only a
few hundred hertz for an air-borne SAR (see Chapter 11), even modest
cross-track velocities can move the scatterer out of the frequency band
used in the processing. The scatterer may then be aliased (or signal may
be lost). Calculating the ensuing displacement effects requires detailed
examination of the way the SAR processing is being carried out.
2 + +
2 +
3 + +
4 + + +
4 + + +
5 + + + +
7 + + + +
11 + + + + +
13 + + + + + + + + +
2~-------------------.rrnnTInr-------------------
t:
'3 0
-1
Figure 6.13 A contour plot of the ambiguity function of the Barker code of length 7 ( T = 1.0,
N = 7).
plateau must be there, to meet the requirements of the total ambiguity having
unit volume.) The cut along the doppler axis is shown in Fig. 6.14b.
A form of phase coding that does not suffer from the restricted length
constraints (and associated sidelobe levels) of the Barker codes is known as
Frank coding 8 • These codes are of length M2, and can be thought of as M
sub-sequences each of length M. Each sub-sequence starts with zero phase.
152 UNDERSTANDING RADAR SYSTEMS
(a)
(b)
wIn
Figure 6.14 Cuts along (a) the delay and (b) the doppler axes of the ambiguity function plot
of Fig. 6.\3.
Code o 0 0 \ I 1 -\ -\ -I o 0 0
'd -6 -5 -4 -3 -2 -\ 0 \ 2 3 4 5 6
Output -\ 0 -\ 0-1 o 7 0 -\ 0 -\ 0-\
The output is obtained by placing a copy of the code, displaced 'd places to the right,
under the code (zero displacement is when the code and its copy are exactly aligned), multiplying
component wise and summing the products.
In the first sub-sequence, all the phases are 0; in the second, the phase of
successive pulses increases by 2n/ M; in the third, it increases by 4n/ M; and
so on (see Fig. 6.15 for the phases of the Frank code of length 16). The
average phase change in the pth sub-sequence is 2n(p - 1)/ M radians. Since
average rate of phase change is a measure of frequency, this means that the
frequency of the coded pulse increases linearly with p. Hence we might expect
the signal to display some of the properties of a linear FM signal. This is
borne out by the contour plot of the ambiguity function of the Frank code
of length 16 shown in Fig. 6.16. The central feature of this plot is a diagonal
DESIGNING RADAR WAVEFORMS 153
;---
I- - r--
I- r-
- ;---
'IT I- r- ;---
T/4
I I Tl2 3T/4 T
Figure 6.15 The phase changes associated with the Frank code of length 16.
1.51-
1.01-
0.5r-
~ 0.01-
()
-0.51-
-1.01-
D
-1.51-
-2.0 -16
I I I oI I I I I
-12 -8 -4 o 4 8 12 16
Figure 6.16 A contour plot of the ambiguity function of the Frank code of length 16 (T = 1.0.
M=4).
154 UNDERSTANDING RADAR SYSTEMS
(a)
1.0
0.5
(b)
whr
Figure 6.17 Cuts along (a) the delay and (b) the doppler axes of the ambiguity function plot
of Fig. 6.16.
ridge similar to that in the ambiguity plot of linear FM (Fig. 6.10). There
are also parallel ridges (such ridges also occur for linear FM, but at a lower
level, and hence are lost by the contour levels used to generate Fig. 6.10).
The cut along the delay axis corresR-0nding to Fig. 6.16 is shown in
Fig. 6.17a. The peak sidelobe level is ../2/16. This demonstrates that the
sidelobe levels of the Frank codes do not fall as rapidly as 1/ M, the rate
achieved by the Barker codes. However, there is in principle no restriction
on the length of a Frank code. For large Frank codes, it can be shown that
the peak sidelobe level declines as 1/(nJ M), so that arbitrarily low sidelobe
levels can be obtained. The cut along the doppler axis is shown in Fig. 6.17b.
6.10 SUMMARY
transmitter is available. Long pulses also permit good doppler accuracy, but,
for simple pulses, give poor range accuracy. The system designer can avoid
this apparent dilemma by using phase-modulated pulses. These can be
constructed to give the large bandwidth that, after pulse compression, leads
to good range accuracy. Whatever transmitted waveform is chosen, the
resolution constraints imposed by the ambiguity function cannot be escaped.
All that the designer can do is to attempt to move the significant areas of
ambiguity into regions of the range-doppler plane where targets are unlikely
to be present.
Key equations
• Effective bandwidth:
<5Wd =
21[(ft
J (2EIN)
• Product of range delay and doppler accuracies:
1 1
<5'd <5eod ~ - -- [J
41[2 EIN
Md = JIPu(tW dt = JIU(wWdeo
2
[sJ
p~(O) 21[E
• Equivalent-rectangle frequency resolution:
L\, =JIPu(wWdw =21[Jlu(tW dt
'" p~(O) E2
156 UNDERSTANDING RADAR SYSTEMS
• Ambiguity function:
6.11 REFERENCES
I. The Fourier Transform and its Applications, R. N. Bracewell, McGraw-Hili, New York, 1986.
2. Probability and Information Theory, with Applications to Radar. P.M. Woodward, Pergamon
Press, Oxford, 1953.
3. A theory of radar information, P.M. Woodward and I. L. Davies, Phil. Mag., 41, 1001,
1950. [Most of this material is covered in reference 2, but not the details leading to the
relation between SNR and accuracy given as Eq. (6.5).]
4. Range and velocity accuracy from radar measurements, R. Manasse, MIT Lincoln Lab.
Report, 312-326, 1955. [This is the original reference, but there is some discussion in
reference 6.]
5. A study of a class of detection waveforms having nearly ideal range-doppler ambiguity
properties, J.P. Costas, Proc.IEEE. 72, 996-1009,1984.
6. Radar Handbook, M. Skolnik, McGraw-Hill, New York, 1970.
7. Group synchronizing of binary digital systems, Barker, R.H., in Communication Theory,
Ed. W. Jackson, Academic Press, New York, pp. 273-287, 1953.
8. Polyphase codes with good nonperiodic correlation properties, R.L. Frank, IEEE Trans.
l'lformation Theory, IT-9, 43-45, 1963.
9. Pulse compression techniques with application to HF probing of the mesosphere, CA.
Gonzales and R.F. Woodman, Radio Sci., 19, 871-877,1984.
6.12 PROBLEMS
6.1 What is the equivalent rectangle resolution of the output of a receiver matched to (a) a
rectangular and (b) a gaussian pulse? What is the frequency resolution of a gaussian pulse?
6.2 Sketch the output of a receiver matched to rectangular pulses of length T [s], if there are
two equal-amplitude targets present separated by range delays of (a) T/4, (b) T/2, (c) T,
(d) 3T/2 and (e) 2T[s].
DESIGNING RADAR WAVEFORMS 157
6:3 Sketch the output from a receiver matched to zero doppler and a rectangular pulse of length
T [s] if there are two targets simultaneously present, both with the same RCS and range but
one of which has zero doppler and one has doppler frequency 2n/T (for simplicity, not reality,
use T= 1).
6.4 Compare the resolutions of the two pulses of problem 4.2.
6.5 A book on radar makes the following statement: 'M pulses, each with duration t p ' can be
viewed as a single pulse of duration Mtp. Since bandwidth B is related to pulselength t by
B = l/t, this combined pulse will have a bandwidth fBI M, where fB is the bandwidth of a
single pulse.'
Consider a radar transmitting a coherent amplitude-modulated pulsetrain u(t) as shown
below.
II (I)
aT
How does the effective bandwidth of this pulsetrain compare with the same quantity for a single
pulse? Show that the amplitude of the Fourier transform of this pulsetrain satisfies
sin(2Tw) I
I
IU(oo)l= sin(Tw/2) IP(w)1
where P(w) is the Fourier transform of a single pulse. Sketch JP(w)l, Isin(2Tw)/sin(Tw/2J1
and 1U (00) I. Do you agree with the statement that heads this problem? Are there any hidden
assumptions in it?
6.6 An air-borne C-band (6 cm) SARtravelling at 300 km h - I carries a I m antenna. It
produces an image in which a ship is apparently displaced by a distance RP/4 m in the
along-track direction at range R, where P is the beam width of the real antenna. What can we
infer about the ship's velocity?
6.7 Find the output from the binary codes:
(a) I, I, I, -I;
(b) 1, I, -I, 1;
( c) I, I, I, - I, 0, 0, 0, 0, I, I, - I, I.
158 UNDERSTANDING RADAR SYSTEMS
This is an example of a complementary code. Sequence (c) is constructed from a pair of shorter
binary codes whose sidelobes cancel each other, and which are separated by (at least) the
length of the shorter codes. Complementary codes give complete cancellation of the range
sidelobes (in the absence of noise) in the vicinity of the main lobe. They are used when very
low sidelobes at zero doppler are important, and are widely exploited in mesosphere-strato-
sphere-troposphere (MST) radars (e.g. Gonzales and Woodman 9 ).
Find all the complementary pairs of length 2, and show that there are no complementary
pairs of length 3.
CHAPTER
SEVEN
SECONDARY SURVEILLANCE RADAR
Secondary surveillance radar is one of the main tools used in air traffic
control.
7.1 INTRODUCTION
Secondary surveillance radar (SSR) is not a true radar system at all but a
two-way communication system between an interrogator on the ground and
transponders fitted to aircraft, which reply automatically. We include SSR
here because the system is very similar to radar in the way it operates, suffers
from many of the classical radar problems and is widely used throughout
the world, often in conjunction with primary surveillance radar.
The origins of SSR lie in the 'identify, friend or foe' (IFF) systems of
World War II; a signal was transmitted from the ground towards a suspect
aircraft, which was required to reply with the appropriate code or be treated
as a foe. Modem SSR works in a similar way; interrogation messages are
transmitted on a one-way uplink frequency of 1030 MHz and cooperating
aircraft reply on a one-way downlink frequency of 1090 MHz. The replies
are fed to a plot extractor, which decodes the aircraft identity and height
IS9
160 UNDERSTANDING RADAR SYSTEMS
and passes them on to the air traffic controllers together with the measured
range and bearing. A classic work on secondary surveillance radar is Stevens!.
Secondary surveillance radar cleverly avoids problems with clutter
through the use of two frequencies, because the receivers at either end of the
link are not in tune with the adjacent transmitter and they do not pick up
unwanted echoes. Another advantage of SSR is that the transmitter and
antenna gain requirements for the uplink and downlink are much more
modest than for a primary surveillance radar operating over the same
range-this is because the combined R2 propagation losses of each one-way
link are much lower than the R4 losses of the two-way radar signal path.
Putting in values:
In practice, the maximum range would probably be less than this because
the aircraft would not necessarily be in the centre ofthe antenna elevation
pattern and receiving the full 52.5 dB W illumination.
In comparison, the power needed by the primary radar can be
worked out using Eq. (2.11) as follows:
(SNR) x N x (4n)3 x R4
PtGt = "2 [W]
Gr x Ii. X (J x Ls
= + 13 [dB] - 145.8 [dB W] + 33 [dB] + 231.6 [dB m4 ]
- (+36 [dB] - 12.7 [dB m 2 ] + 20 [dB m 2 ] - 5 [dB])
= 93.5 dB W
Assuming 36 dB as the antenna contribution to the ERP, then the
transmitter power would need to be over 500 k W.
Primary and secondary radar are often used together (sometimes the
antennas are even attached and rotate together); the primary system is used
to provide air traffic controllers (A TC) with a 'map' on a plan position
indicator (PPI) display of everything moving in the sky in their region. The
SSR interrogates each target, usually with the request: 'Who are you and
what height are you at?' Cooperative targets, such as all civil airliners and
most private and military planes, reply with the information requested, which
is then displayed in alphanumeric form at the appropriate place on the PPI
display. In today's crowded airlanes, this information is valuable, for after
an aircrew have filed their flight plans, including the flight identification
number, the aircraft can be tracked automatically along its route without
162 UNDERSTANDING RADAR SYSTEMS
the need for ATC requests for the plane to identify itself. Such requests would
be necessary if the only information were from primary radar.
Identify 3/A
Height? C
Aircraft reply to this interrogation with a train of pulses that are 0.45 its
wide and spaced 1.45 its apart and which are used to modulate the 1090 MHz
downlink carrier. The first and last pulses in this train are always present
and are known as framing pulses F 1 and F 2 (Fig. 7.2). In between F 1 and
F 2 are 12 pulses, which mayor may not be present, depending on the message
AmPlitudetJ
F1LJP4'--________________---I1F1L
Time
Figure 7.1 The interrogation message is determined by the separation of the pulses PI and
Pl' Pulse P2 is used for sidelobe suppression.
Amplitude
i F,
Figure 7.2 The form of the reply from the transponder on the aircraft.
SECONDARY SURVEILLANCE RADAR 163
being transmitted, and an extra centre pulse, which is not currently used.
The 12 pulses are used as a 12-bit code, which has 4096 possible combinations
-enough to give the aircraft identity when in mode 31 A. In mode C, one
of the pulses is not used, but the remaining 2048 codes are sufficient to give
the aircraft height in steps of 100 feet. When both height and identity are
required, the requests are made alternately, which is known as mode
interlacing. There is plenty of time to interlace modes A and C because, as
the SSR antenna rotates, the beam illuminates an aircraft for about 30 ms
and during this time its transponder is interrogated about 15 times.
Normally, an airline pilot will select mode A/C and the transponder is
programmed to respond to requests from the ground without crew inter-
vention. The aircraft altimeter automatically feeds height information to the
transponder. There are also four other switches the aircrew can set, which
send out additional messages for unusual situations such as a hijacking,
radio communications failure or if ATC are having problems identifying the
aircraft.
Secondary surveillance radar is used to locate aeroplanes in a similar
way to primary radar. Ranges are measured by the round-trip time of a
pulse travelling up to the aircraft and returning, with appropriate allowances
for the delays in the equipment on the aircraft and on the ground. The
accuracy of the range measurement can be improved by using correlation
methods on the entire received pulsetrain, rather than using just one pulse.
The aircraft height is known from the information given by its own altimeter,
and so the remaining piece of information required to fix its position is the
azimuth, as measured at the SSR.
Azimuths are now measured using the monopulse technique described
in Chapter 3, and this provides much greater accuracy than earlier methods.
The importance of monopulse for SSR is that it provides a measurement of
angle on every pulse, so that when a train of pulses is received, these
measurements can be combined to form an improved estimate of the azimuth
(in general, n independent measurements can be combined to improve the
J
accuracy by a factor of n). A typical SSR antenna is about 8 m wide,
giving an azimuth beam width of about 2.5°, but after monopulse processing
the error in the azimuth of the aircraft is as low as a few minutes of arc.
Because the antenna is not required to measure elevation angles (the aircraft
height being already known), the avertical size of older SSR antennas
has often been only about 0.4 m, giving vertical beamwidth approaching
50°. These long, thin parabolas have given rise to the descriptive term
'hogtroughs '.
Modern SSR antennas are usually flat phased arrays having vertical
dimensions of the order of 1.6 m to give improved control over the elevation
pattern, and these have become known as LV As, large vertical apertures.
However, they remain several times wider than they are tall, and to some
extent still retain the long, thin appearance of the old hogtroughs.
164 UNDERSTANDING RADAR SYSTEMS
~AZimuth
rotation
Pattern of
main SSR antenna
/ '" '"
4::i::
/ '"
/ '"
/
/ '"
/
/
/
",-
" , '" /
/ 'K
/ /"'\
I \
I \
I ~ Pattern of omnidirectional
\ I
\ antenna
\ I
\ /
\ /
"- /
/
" '- _/
Figure 7.3 SSR main-beam omnidirectional antenna patterns. Aircraft at close range may be
interrogated by a sidelobe of the antenna as well as by the main beam.
SECONDARY SURVEILLANCE RADAR 165
P 2 exceeds PI or P 3' then the request must have come from a sidelobe of
the main antenna and the transponder makes no reply.
The idea of using a control beam is an example of good engineering.
Rather than spending a great deal of money trying to improve sidelobe
performance of the main antenna, the problem can be solved by adding an
extra low-budget antenna. Most modern SSR systems do not in fact have a
separate control antenna-the control beam is synthesized by the main
antenna, so that both main beam and control beam suffer equal multipath
fading (see next section), and amplitude comparisons between the two remain
valid.
Other problems and solutions involving SSR systems are as follows:
The problem of radio signals arriving at the receiver by more than one route
and corrupting the information carried is common to most forms of radar
and to radio communications in general. It is usually known as multipath
and frequently involves reflections from the ground as well as buildings.
166 UNDERSTANDING RADAR SYSTEMS
7.4 MULTIPATH
Figure 7.4 shows a radio signal travelling from a radar system on a tower
up to an aircraft; some of the transmitted energy travels by the shortest path
and some is reflected from the ground and appears to come from a source
beneath the ground. These two radio sources act in the same way as Young's
slits in optics, and give sum and difference interference patterns. In fact, a
version of the Young's slit experiment using a single source of light and a
mirror to generate an apparent second source is called Lloyd's mirror, and
is an even better analogy of the multipath case.
When reflected from the ground, the radio signal undergoes a phase
change, which is nominally 180 If the difference in path length between the
0
•
Figure 7.4 Propagation over a plane reflecting surface results in two signal paths that interfere
with one another.
SECONDARY SURVEILLANCE RADAR 167
\
"-
"-
"-
"-
"-
"-
j
""- "-,
,,
,,
,,
\
Figure 7.5 The presence of a reflecting surface causes the antenna pattern to break up into
lobes at low elevation angles.
The process of achieving this has already begun with the new LV A antennas,
which have a much better control over the vertical beam pattern. Other
techniques that might be considered are:
For all its association with the world of radar, SSR is essentially a digital
communications system, and a new mode of operation known as mode S is
coming into service to exploit this aspect more fully, see Scanlan 3 . Longer
pulsetrains are transmitted to increase the information transfer, and the
interrogation of aircraft is selective in that the ground base addresses them
one at a time.
The first task of a new SSR system is to find out which of the aircraft
it is interrogating are equipped with mode S transponders. This identification
is achieved by the use of a fourth interrogation pulse P4, which can be given
a duration of either 0.8 or 1.6 J.LS. Because pulses PI to P 3 are transmitted
as before, the older mode Ale transponders continue to reply as normal. A
mode S transponder examines the duration of P 4; if it is set to 0.8 J.LS the
transponder makes no reply, but if P 4 = 1.6 J.LS then it replies by giving its
own unique address. These addresses are noted by the SSR system, which
then schedules the mode S transponders for special interrogation.
The mode S interrogation proper is contained within a long P 6 pulse,
which is modulated by differential phaseshift keying (DPSK-a standard
digital communications technique) to convey either a 56- or 112-bit word
depending on whether the length of P 6 has been set to 16.25 or 30.25 J.LS.
SECONDARY SURVEILLANCE RADAR 169
These bits are used for selectively addressing the mode S transponders.
Aircraft fitted with the older transponders do not reply to this new type of
transmission because, cleverly, the first two pulses PI and P 2 are set to the
same amplitude, which older transponders treat as a sidelobe interrogation
and .therefore ignore.
Mode S transponders reply in a similar manner to the interrogation; a
four-pulse preamble is followed by a 56- or 112-bit binary word. Several of
these words may be strung together in successive transmissions to enable
quite complex messages to flow between aircraft and the ground.
7.6 SUMMARY
7.7 REFERENCES
1. Secondary Surveillance Radar, M.C. Stevens, Artech House, Norwood, MA, 1988. [The
classic book on modern SSR; it is clear and full of detail.]
2. Understanding Radar, H.W. Cole, BSP Professional Books, Oxford, 1985. [Contains more
than 60 pages of information on SSR in a very readable form.]
3. Modern Radar Techniques, Ed. MJ.B. Scanlan, Collins, Glasgow, 1987. [Chapter 6, also
written by H.W. Cole, is dedicated to modern SSR.]
7.8 PROBLEMS
7.1 The Cossor Condor 9600 is a complete ATC system that includes full monopulse SSR. One
of the SSR antenna options is the Condor 9642 large vertical aperture, which has a peak gain
of 27 dB and a beam width of 2.45 at the - 3 dB points.
0
If the power of the Cossor interrogation transmitter is adjusted so that the ERP conforms
to the ICAO specification of 52.5 dB W when it is used in conjunction with the 9642 antenna,
would you expect an aircraft transponder to receive an adequate signal at the maximum
instrumented range of 256 nautical miles? Assume additional propagation losses of 2 dB and
an antenna pattern loss of 3.5 dB because the elevation angle of the aircraft places it above the
angle of maximum antenna gain.
170 UNDERSTANDING RADAR SYSTEMS
7.2 For the downlink the ICAO defines the transponder output power as 24 dB W. For an
aircraft at a range of 100 nautical miles, what signal strength would the Condor 9600 system
receive? Assume 1 dB loss for atmospheric attenuation and a further 7 dB for system and vertical
antenna pattern losses. Remember that the downlink frequency is not the same as the uplink
frequency used in the previous question.
7.3 If the minimum working signal level of the interrogation system on the ground is
- 110 dB W, what is the theoretical maximum downlink range of the Condor system? Allow
3 dB in your calculations for azimuth beam width loss.
7.4 Why does the downlink apparently give better performance than the uplink if the transmitter
is less powerful?
7.5 Estimate roughly the angular uncertainty that might be expected when the Condor system
tracks an aircraft with a range of 100 nautical miles. Assume a receiver noise level of - 130 dB W.
CHAPTER
EIGHT
PROPAGATION ASPECTS
How to cope with radio waves not travelling In straight lines in the
atmosphere.
8.1 INTRODUCTION
Radars frequently operate through the atmosphere of the earth, often at low
elevation angles where most targets occur, but where there are most problems
for the radar. Aircraft flying at constant altitude towards a surveillance radar
appear over the curvature of the earth and are difficult to detect and track.
Similar problems occur with ship surveillance when using maritime radar.
When close to the horizontal, radar beams have their greatest path length
through the atmosphere, which is itself at its most dense and turbulent.
Low-elevation radar beams can also encounter obstacles, such as hills, and
become diffracted into the shadow regions behind.
In this chapter we are going to investigate the radio horizon and discover
whether the atmosphere, the ionosphere and the terrain significantly affect
the propagation of radio waves. Again, elementary optics is useful to describe
what is happening.
171
172 UNDERSTANDING RADAR SYSTEMS
Assuming for the time being that radio waves travel in straight lines through
the atmosphere, the radio horizon is defined by a line tangent to the surface
of the earth, as in Fig. 8.1. If R is the range of a point on the radio horizon
(or of a target appearing at zero elevation), h is the height of the point and
Re the radius of the earth, then
R2 = (Re + h)2 - R;
= 2Reh - h 2 [m 2 ]
It is reasonable to assume that h 2 « 2Reh and therefore
R ~ j(2Reh) Em] (8.1 )
Although the earth is not truly spherical, being flattened at the poles and
bulging slightly at the equator, there is little error in assuming it to be a
Centre of
Earth
Figure 8.1 Geometry for calculating the height of the radar horizon as a function of range.
PROPAGATION ASPECTS 173
sphere with an average value of Re = 6378 km. Many radar engineers have
to work with aircraft height in feet and range in nautical miles. (A nautical
mile is the distance around the earth's surface corresponding to an angle
of one minute measured at the centre of the earth. It equals about 6080 feet
or 1.85 km; see Appendix III.) Restating Eq. (8.1) in nautical miles
[no mile], but with h in feet, gives
R ~ )(2 x 3444h/6080) [no mile]
and a rough approximation that can prove useful is
Rn. mile -- ) hreet [no mile] (8.2 )
If the radar is at an altitude hradar> Eq. (8.1) can be modified as
R ~ ) (2Rehradar) + ) (2Rehtarget ) em] (8.3 )
SOLUTION Using Eq. (8.1) and Re = 6378 km, we get the following
answers: Sea-skimming missile is detected at a range of 11.3 km.
Low-flying fighter is detected at a range of 35.7 km. High-flying bomber
is detected at a range of 437 km.
...'" 100
'"0-
..c Water vapour resonance
<J> - 22 GHz
0
E
co
'"
.::
C
10
'"'"
-=
..c
bI)
::l
2
-=c
.2 resonance
co::l - 60GHz
c
~
co
>.
<II
~
C
~
f- 0.1
0.1 10 100
Radar frequency [GHz)
Figure 8.2 Two-way atmospheric absorption through the entire troposphere as a function of
radio frequency for two elevation angles.
p= ( ----.!2cos¢
. ndh
1d )-1 Em]
Figure 8.3 Atmospheric refraction causes the apparent height of a target to be greater than
its true height.
176 UNDERSTANDING RADAR SYSTEMS
radius of the earth has the effect of bending the radio waves back up into
a straight line over this artificial earth, and allows us to return to our simple
geometry provided we remember to replace Re by kRe every time it occurs
in the calculations. But what value should we choose for k? Geometrically
k is given by
k = p/(p - Re) [ ] (8.8 )
Combining this transform with Eqs (8.4) and (8.6), and assuming cos 4> == 1
for low elevation angles, gives
k = ( I + 10- 9 Re : : ) - 1 [ ] (8.9)
140000~ 40 r-----..j_1
\ Larger targets
detectable
120000~
_ 100000
v E 30 r--+--ic....LJ
~ ::!.
80000
~
'<; 20 r---ftf+-I-LI
:I:
40000
10
oL~~~~~~~i:=~~:~;:t:=~joo
20000
Figure 8.4 A typical graph for plotting radar vertical coverage, using the four-thirds earth
model.
PROPAGATION ASPECTS 177
SOLUTION Using Eq. (S.1), but with Re replaced by 1Re we now have:
Sea-skimming missile is detected at a range of 13.0 km. Low-flying fighter
is detected at a range of 41.2 km. High-flying bomber is detected at a
range of 505 km.
/'
/'
/'
/'
,-
,-
,-
,-
,-
,-
/'
~
Figure 8.S Super-refracting ducts can cause radar signals to bend back to earth, creating
anomalous clutter problems.
178 UNDERSTANDING RADAR SYSTEMS
em] (8.12 )
-----------4.~----------d2----------~.1
Figure 8.6 The geometry used to evaluate diffraction effects.
terrain in terms of these obstacles. Fortunately, most hills and ridges can be
described quite well by single and multiple diffraction edges or by cylinders
with knife-edges on top of them. Procedures exist for determining the outcome
when there are several different types of these obstacles in the path of the
radio wave.
The simplest and most rigorously analysed diffraction obstacle is the
knife-edge, shown in Fig. 8.6. Various texts describe the mathematics; see
for example Griffiths 5 or Meeks 1 (which includes useful program listings).
Although a full mathematical analysis is complicated, the results may be
summarized, using the notation shown in Fig. 8.6, as follows:
Radars used on the battlefield for the detection or guidance of weapons are
interesting in many respects, not least of which is the sophisticated command
ISO UNDERSTANDING RADAR SYSTEMS
Retardation
Measurement of
Target
errors
I
~r
300
E
.=..,
"0
oE 200
<:
Figure 8.8 The variation of electron density with altitude in the day-side ionosphere.
III =
f
s
n ds ~ 40.3
-2
j
f
s
Ne ds [mJ (8.15 )
where s is distance along the path S from the radar to the target. The integral
of electron density along the path is called the total electron content (TEe),
for obvious reasons. It varies considerably with position and time, but
generally does not exceed 5 x 10 17 electrons. This value of TEe would give
rise to a one-way range error of 900 m at 150 MHz, which is the frequency
used, for example, by the TRANSIT positioning system. More problem-
atically, variations in TEe would cause this error to vary between a few
metres and hundreds of metres unless they were properly corrected.
(a)
True Apparent
position position
(b)
Refracting
slab
""
""
""
""
"
True " Apparent
position position
i i o
Planar
phase fronts
Irregular
ionosophere
Perturbed
phase
front
Propagation
to
ground
Amplitude
scintillations
Figure 8.10 The mechanism by which phase perturbations in the ionosphere give rise to
scintillations observed on the ground. As the perturbed wave propagates to the ground,
interference effects develop, and give rise to large fluctuations in phase and amplitude.
hence known as scintillations. On the return path to the satellite, the same
type of disturbance will occur.
The amplitude scintillations can be thought of as fluctuations in the ReS
of the target, and hence are treated in the same way as for the Swerling cases
(though there is still no universal agreement on the correct PDF to describe
these scintillations). Whether the received signal varies on a pulse-to-pulse
or a scan-to-scan basis depends on the relative motion of the satellite, the
target and the ionosphere. Phase scintillations can also have important effects
on systems that rely on phase coherence, such as space-borne SARs. If the
pulse-to-pulse signal contains a random phase element that changes signifi-
cantly as the synthetic aperture is being formed, the radar performance may
be degraded. Progressively more serious effects include displacement of the
beam (leading to geometric errors), increased sidelobe levels (leading to loss
of contrast in the image) and destruction of the focus (leading to complete
loss of the image). Figure 8.11 shows all these effects for simulated SAR data
as the operating wavelength increases. At a wavelength of 0.83 cm,lthe synthetic
PROPAGATION ASPECTS 187
Distance [m I Distance [m I
5.0 5.0
-50~.0~~~~~O~~~~~5~0.0 -50.0 o 50.0
A =0.83cm
-45.0 ~ -45.0
I a:l ~
~ ~
c c
'<;j '<;j
o o
Distance [m I Distance [m)
5.0 5.0
-50.0 o -50.0 o 50.0
A = 6.00cm A = 24.00cm
-45.0 ~ Cil
~ ~
c -45.0 c
'<;j '<;j
0 0
Figure 8.11 Simulated distortions of the synthetic aperture gain pattern caused by ionospheric
irregularities. The different patterns correspond to increasing wavelength (and hence increasing
synthetic aperture length; see Chapter 11), as marked on each pattern. The ionosphere was
the same in all cases.
beam is hardly disturbed. By the time the wavelength has increased to 24 cm,
the beam has been almost completely destroyed.
8.7 SUMMARY
Key equations
• Range of a point on radio horizon:
R ~ J(2R eh) [mJ
which can be restated in nautical miles as:
[no mile]
• Radio refractivity:
N = (n - 1) x 106 [J
which can be calculated from:
k = ( 1 + 10 - 9 Re ~~) - 1 [ ]
• Faraday rotation:
n=M x (TEC)/P [radians]
PROPAGATION ASPECTS 189
8.8 REFERENCES
There are many good books on the subject of propagation because this is a
subject of importance in communications as well as in radar. The following
selection is suggested as a useful introduction to the subject.
I. Radar Propagation at Low Altitudes, M. L. Meeks, Artech House, MA, 1982. [A very
readable monograph complete with program listings and a good bibliography.]
2. Radio Meteorology, B. R. Bean and E. J. Dutton, Dover Publications, New York, 1968.
[This is the standard text on large-scale atmospheric variations in radio refractive index.]
3. Radiowave Propagation, Ed. M. P. M. Hall and L. W. Barclay, Peter Peregrinus for the
IEE, Stevenage, Herts, 1989. [Originally derived from course notes accompanying an lEE
school on radio wave propagation, this book forms an indispensable guide to the whole
subject.]
4. Reference Atmosphere for Refraction, Recommendations and Reports of the CCIR, CCIR
Rec369-3, V, lTV, Geneva, 1986.
5. Radio Wave Propagation and Antennas, J. Griffiths, Prentice-Hall, Englewood Cliffs, NJ,
1987. [A particularly easy to read guide to propagation.]
6. Radar Handbook, Ed. M. I. Skolnik, McGraw-Hili, New York, 1970.
7. Radio Wave Propagation, A. Picquenard, Macmillan, London, 1974.
8. Modern Radar Techniques, Ed. M. J. B. Scanlan, Collins, Glasgow, 1987.
8.9 PROBLEMS
8.1 At what range could a target flying at a height of I km be detected by a radar at sea level,
assuming (a) no refraction, (b) four-thirds Earth, (c) a refractive gradient of - 20 N jkm and
(d) a refractive gradient of - 120 N jkm ?
8.2 In the cases (c) and (d) in problem 8.1, what range corrections would be needed to allow
for atmospheric retardation? Assume No = 300 at the radar site.
8.3 (a) In the next chapter it is shown that the power received by a weather radar, when
precipitation fills the beam, is proportional to 1j R2. If the radar were badly sited such that
many of the low-elevation observations were made at grazing incidence to the local horizon,
how would the performance be impaired when scanning for light rain nominally detectable at
4.5 km in the clear?
(b) How would the detection range of a point target be affected at the same elevation?
8.4 A space-based radar operating on the two frequencies 200 and 400 MHz measures the range
delay to a target as 44 and 14 ms respectively. What is the true range to the target?
CHAPTER
NINE
RADAR STUDIES OF THE ATMOSPHERE
9.1 INTRODUCTION
Hazard avoidance Most airliners carry a weather radar in the nose to look
ahead and give a warning of severe weather on a display inside the cockpit.
Pilots may then take avoiding action within the limits set by their flight
lane. Airports prone to severe storms are now also installing ground-based
weather radars to give warnings of severe down-draughts, which are a
danger to aircraft taking off and landing. At present, strong down-draughts
during thunderstorms cause about one aircraft crash every 18 months in
the USA alone.
190
RADAR STUDIES OF THE ATMOSPHERE 191
Hydrometeors
This is the meteorological term for scattering particles (rain, ice particles,
etc.). They cause the Rayleigh scattering that we encountered in Chapter 2
because the radar wavelength A is usually greater than the diameter D of the
particles. The ReS of a single droplet is
= C n 5D /-A.
6 4
(J particle (9.1 )
where C is a dimensionless constant that depends on the dielectric constant
of the particles. For water, C is near 1, whereas for ice particles it is about
0.2. To calculate the RCS of a rain cloud, we must sum over the volume
resolution cell of the radar, to give
(9.2 )
Precipitation a b
and the radar equation becomes (with any appropriate units also shown)
[W] [ ][ ][ ][ ] [m 3 ] [rad][rad][m S-I ][s][ ) [ )
r-A--.
[W] PI GI Gr C n2 (to- 1H Z) de drj> c r L,ys Latmos
[W]
Pr 64 ).2 R2 2
[ ][m 2 ][m 2 ] [ ]
(9.6 )
Horizon
0--'-/-+--- Ship
Aircraft travelling
' \ - - - Backscatter from the atmosphere radially towards
the radar
Figure 9.1 Backscatter from rain and other atmospheric features amount to a form of moving
clutter that can mask slowly moving targets.
the pulse duration is 1.0 p.s, what echo power would you expect to receive
from heavy rain (r = 4 mm h - 1), if the rain entirely filled a resolution
cell, at a range of 50 km?
Z= 32.6dB mm 6 m- 3
AO A<b = -47.2 dB rad 2
r= -60.0dB s
l/A? = 20.2 dB m- 2
1/R2 = -94.0 dB m- 2 (Range = 50 km)
Pr = -80.9dB W
The received power is well above the expected noise level for a I MHz
bandwidth, even allowing for additional system and propagation losses.
Angels
This intriguing name was given to the echoes of discrete targets that returned
from an apparently empty sky to trouble early radar operators. Birds are
probably the main cause, but clouds of insects can also cause angels and,
because they drift in the wind, they can cause clutter problems similar to
rain clouds.
The RCS of a bird generally lies in the resonance region and large
variations are observed, but a rough rule is that a bird has the same echoing
area as a plastic bag filled with water having the same weight. Typical RCS
values are -40 dB m 2 for a small bird and -20 dB m 2 for a larger bird such
as a gull. Flocks of birds appear as much larger, moving targets. Radar can
be used to study bird migration, and to identify birds by wing beat frequency,
altitude and speed; see for example Eastwood 6 .
Insects have an RCS of - 50 dB m 2 or lower, but, again, concentrations
of insects within the volume resolution cell of the radar can cause sufficient
scattering to create the angel effect.
196 UNDERSTANDING RADAR SYSTEMS
Clear-Air Turbulence
Echoes from genuinely clear air were originally quite unexpected and difficult
to explain, although it was guessed that weak fluctuations in the radio
refractive index of the air must be responsible. It is now generally understood
that echoes from clear air arise from two main types of scattering mechanism:
one is volume scattering from turbulence and the other is specular (mirror-
like) reflections from thin layers.
The atmosphere is everywhere turbulent but there exist regions of greater
intensity such as convective cells or 'thermals'. Volume scattering from these
regions (and also from the whole atmosphere, but fainter) is due to
constructive interference from fluctuations in refractive index. Those scatterers
with a scale size equal to half the radar wavelength are mainly responsible
for the backscattering, in a process similar to the Bragg scattering discussed
in more detail in Chapter 11.
One of the hazards of flying is to encounter the sudden turbulence that
forms in thin horizontally stratified layers and which sometimes has a mean
vertical gradient, or refractive index, much greater than in the surrounding
atmosphere. These layers may be only a few tens of metres in thickness and
yet they can extend horizontally for tens of kilometres. The radio scattering
from these layers causes strong radar echoes in both the troposphere and
the lower stratosphere. The stratosphere is the horizontally stable layer above
the troposphere and extends in altitude from roughly 12 to 50 km.
The region above the stratosphere is called the mesosphere, and thin
layers are found here too. The early Soviet cosmonauts produced some
excellent free-hand sketches of mesopheric layers that are visible when viewed
edge-ways on against the bright limb of the earth. They have since been
confirmed by radar observation. Above 100 km, the atmosphere can no
longer be considered entirely neutral, and ionospheric scattering processes
dominate.
The refractive index of the neutral atmosphere was described by
Eq. (8.5), but in the mesosphere there are sometimes free electrons present,
which cause an ionospheric term to be added:
[Hz]. Equation (9.8) describes the refractive index n, from ground level up
to 1()() km (to get the radio refractivity N, subtract 1 and multiply by 106 ).
Using reasonable values of T, e, p and Ne we find that the wet term (containing
e, the partial pressure of water vapour) dominates in the troposphere but
above 50 km the ionospheric term begins to determine the refractive index.
The usual practice in atmospheric physics is to incorporate refractive index
RADAR STUDIES OF THE ATMOSPHERE 197
changes into the radar equation by adopting yet another version, based on
Eq. (9.6):
p = P)2LSYs(C; + C;) [W] (9.9 )
r 16n 2 R2
9.3 MESOSPHERE-STRATOSPHERE-TROPOSPHERE
RADAR
During the 1970s it was realized that VHF radars operating at wavelengths
of 1-10 m had considerable potential for studying atmospheric dynamics
because these wavelengths are well matched to turbulence scale sizes. These
radars, known as mesosphere-stratosphere-troposphere (MST) radars,
have several advantages over rocket, balloon and aircraft measurements,
including:
100m
Figure 9.2 Typical MST radar arrangement using several fixed beam positions to measure
horizontal and vertical air motions.
RADAR STUDIES OF THE ATMOSPHERE 199
Pr = -165.8 dB W
If the system were internally noise-limited, the level would be
-204 dB W Hz- I + 60dB (1 MHz bandwidth) plus the noise factor
of the receiver. In this case the volume scattering echo from a single
pulse would be below noise level. In practice, a coherent integration
time of about 1 s would be used and the final bandwidth would
therefore be I Hz. However, external noise dominates internal noise by
about 19 dB at this frequency, and so the final noise level is about
-204 + 19 = 185 dB W, and volume scattering would be detected with
a 19dB SNR.
COMMENTS Roughly speaking, at 11 km the signal strength falls by about
1 dB per kilometre because of the 1/ R2 factor, and a further 1 dB per
kilometre because of the change in C~, giving a total of 2 dB per
kilometre. If no useful atmospheric science can be carried out with a
SNR of less than 10 dB, these results imply that the radar would be a
useful instrument only up to an altitude of about IS or 16 km (longer
coherent integration cannot be used owing to the short timescales of
turbulent processes). The presence of any scattering layers, adding a C;
term into the radar equation, would increase the SNR at lower altitudes,
but by 20 km, they too are beginning to disappear. There is no escape
from the conclusion that higher-altitude atmospheric research requires
more powerful transmitters, larger antenna farms and more money.
200 UNDERSTANDING RADAR SYSTEMS
A MST radar will measure returns from the troposphere and the lower
stratosphere, and may sometimes receive echoes from the mesosphere when
free-electron scattering occurs, but there is no easy way of investigating the
region of the middle atmosphere between the two. For this reason, some
systems were not designed to try to see beyond the lower stratosphere and,
as a result, two distinct types of radar have evolved: the true MST system,
and the smaller stratosphere-troposphere (ST) radars.
An example of a powerful modern MST radar is the Japanese MU radar,
mentioned in problem 9.6 at the end of the chapter. The MU system uses
an active antenna array (see Chapter 15) consisting of 475 solid-state
transmitter modules, and Vagi antennas, to develop a total peak transmitted
power of 1 MW. At the other end ofthe scale, there is an increasing use ofST
radars to study turbulence and severe wind shears in the lower atmosphere,
and in the USA, operational ST systems are to be installed at 47 of the
country's busiest airports to warn pilots of potentially dangerous wind
conditions. Some of the lower-atmosphere radars are more compact UHF
wind profilers, which operate in much the same way, but at frequencies around
400 or 900 MHz.
New methods of operating ST and MST radars, and of carrying out the
signal and data analysis, are continually being developed, and this field of
radar is expanding rapidly.
Winds in the upper atmosphere between about 70 and 110 km have been
studied since the 1950s by meteor wind radar. When tiny meteoritic particles
enter the earth's atmosphere, they heat up due to friction (like the Space
Shuttle during re-entry). The heating causes a meteor to boil away in a
process known as ablation, which leaves behind a long, slightly cone-shaped,
column of ionization, which is a strong scatterer of lower-frequency (HF
and VHF) radio waves (see Fig. 9.3). The ionized trails drift in the neutral
wind with a velocity that is easily measured by pulse doppler radar. There
are always sufficient numbers of sporadic meteor echoes per hour for even
a radar of modest power to be able to measure the larger-scale atmospheric
features. At regular times each year, the earth passes through comet debris
causing meteor showers that give an increased echo rate for a few days.
Although meteor wind radars do not have the resolution of the newer
MST radars, they remain of considerable interest because the meteor trail
formation process gives some clues about upper-atmosphere chemistry
(recent work suggests that this could also include ozone concentrations).
There is also an interest in over-the-horizon communications by forward
scatter from meteor trails, and astronomers are interested in learning more
about the distribution and sizes of meteoritic material within the solar system.
RADAR STUDIES OF THE ATMOSPHERE 201
Figure 9.3 Radar scattering from meteor trails can be used to measure winds at high altitudes.
where rj = initial radius of the trail em], which has been found empirically
to be about 0.22 m at 80 km altitude, O.S m at 90 km and 1.1 m at 100 km.
Using 0.5 m for the initial trail radius gives (12 = 0.7.
The third term (13' and the exponential factor, allow for the expansion
of the trail by diffusion; (13 is given by
[ ] (9.16)
The constant K [m] is related to the way the trail diffuses and the velocity
of the meteoritic particle. The value of K is about 0.09 m at 80 km, 0.05 m
at 90 km and 0.1 m at 100 km. A typical value for (13 is therefore 0.6.
The exponential term gives a characteristic decay to meteor echoes,
which is used in some systems as a method of determining the echo height.
The time constant for the received power to decay by a factor e 2 is given by
[s] (9.17)
best known, and most spectacular, meteor showers are the Perseids, around
12 August each year, and the Geminids, around 13 December.
Besides their use in atmospheric physics, these formulae describing
meteor rates and ReS can be used to estimate the extent of meteor clutter
detected by various types of radar (for example, HF radars, described in the
next chapter, experience meteor clutter problems).
9.6 SUMMARY
From initially being a nuisance, the scattering of radar signals from the
atmosphere has been turned into a useful, and expanding, research technique
for the study of atmospheric physics and meteorology. Scattering occurs
from discrete sources (rain, birds, etc.) and also from changes in the refractive
index of the air, mainly caused by turbulence.
Weather radars, investigating lower-atmosphere cloud physics and
precipitation, operate at frequencies in S-band and above, but new VHF
phased array radars have emerged as a tool for probing the atmosphere from
1 to 100 km altitude.
204 UNDERSTANDING RADAR SYSTEMS
Key equations
• Radar reflectivity factor:
Z = arb
• Radar cross-section of resolution cell:
s
(J = Cn ZR 2 /1() /1</> CT
4
..1. 2
• Radar equation for hydrometeor scattering:
p = 2.3 x 10- 11 PtGtGrZ /1() /1</> rLsysLatmos
r A2R2 [W]
9.7 REFERENCES
I. Radar in Meteorology, Ed. D. Atlas, Battan Memorial and 40th Anniversary Radar
Meteorology Conference, American Meteorology Society, 1990. [A large but fascinating
book full of the history of meteorological radars, as well as current thinking.]
2. Radar Meteorology, L. J. Battan, University of Chicago Press, Chicago, 1959.
3. Radar Design Principles, F. E. Nathanson, McGraw-Hill, New York, 1969.
4. Introduction to Radar Systems, M. I. Skolnik, McGraw-Hill, New York, 1985.
5. Radar Propagation at Low Altitudes, M. L. Meeks, Artech House, MA, 1982.
6. Radar Ornithology, Sir. E. Eastwood, Methuen, London, 1967.
7. Meteor Science and Engineering, D. W. R. McKinley, McGraw-Hill, New York, 1961.
8. Turbulence at altitudes of (80-100) km and its effects on long duration meteor echoes,
J. S. Greenhow and E. L. Neufeld, J. Atmos. Terr. Ph),s., 384-392, 1959.
9. Radio propagation by reflection from meteor trails, G. R. Sugar, Proc. IEEE,52, 116-136,
1964.
10. Laser Monitoring of the Atmosphere, Ed. E. D. Hinkley, Springer-Verlag, Berlin, 1976.
9.8 PROBLEMS
9.1 The UK Meteorological Office weather radar network has used Siemens-Plessey S-band
and C-band radar sensors to measure rainfall. The Siemens-Plessey type 45S sensor features
a 3.66 m diameter parabolic antenna giving a 2° pencil beam and a gain of 37 dB at 2.9 GHz.
The peak transmitter power is 650 kW and the pulse duration is 21ls. If the noise level is
-137 dB W, what is the SNR of the echo from a rain cloud filling the beam at a range of
20 km, if the rainfall rate is 1 mm h - 1 ?
9.2 What would be the effect of increasing the rainfall rate to 100 mm h - 1, in the above problem?
9.3 What would be the dynamic range requirement of a receiver covering rainfall rates of 1;8
to 64 mm h - 1 and ranges from I to 200 km? Could this be reduced by sweeping the gain of
the receiver to increase the sensitivity with increasing range? If so, what power law should be
used for the swept gain?
9.4 The German SOUSY MST radar system located in the Harz Mountains operates at
53.3 MHz and has an antenna array 70 m in diameter giving a beamwidth of 5' and a gain of
31 dB. The beamformer uses 4-bit phaseshifters (see Chapter 15) to steer the beam anywhere
within a 30' cone centred on the vertical. At 7' off-vertical, it is found that the echo power
received from a thin layer at an altitude of II km is 10 dB lower than the measurement in the
vertical beam. Could this be explained by the geometry and the longer range to the layer in
the off-vertical beam?
9.5 The peak transmitter power of the SOUSY radar is 600 kW. What is the strength of an
echo scattered from turbulence at II km, assuming a 21ls pulse duration and 7 dB system losses?
9.6 The Japanese MU MST radar system operates at 46.5 MHz, uses a circular antenna array
103 m in diameter and a peak transmitter power of I MW. How much larger would the SNR be
than that of the SOUSY system, assuming the other parameters are identical?
9.7 If the UK MST radar (near Aberystwyth, Wales) uses a coherent integration of 512 pulses.
and is set to have an unambiguous range of 24 km and a total observation time of 10.5 s, what
is the width of the doppler spectrum in the fast Fourier transform?
9.8 If the UK MST radar uses a vertical beam, and beams steered to a maximum of 12'
off-vertical, would an aircraft travelling at 300 m S-1 cause any problems?
9.9 Can an MST radar be used as a meteor radar? Would you expect to detect many meteors
in the vertical beam?
CHAPTER
TEN
OVER-THE-HORIZON RADAR
Over-the-horizon radars are even less perfect than microwave systems, but
the rewards for seeing over the horizon are worth pursuing.
10.1 INTRODUCTION
It has been known since the experiments of Guglielmo Marconi in 1901 that
radio waves could propagate beyond the horizon because of the signals he
successfully transmitted from Poldhu in Cornwall, UK, to St John's,
Newfoundland. Investigations into the cause of this propagation soon
revealed that the solution to Maxwell's equations for a wave at a plane
interface between two media gives a space wave (free-space propagation)
and a surface wave (a wave guided along the interface). With the discovery
of the earth's ionosphere in the 1920s, it was realized that there was also a
third possible mode of propagation, the ionospheric wave, which turned out
to be the explanation of Marconi's transatlantic communications.
The propagation of HF (3-30 MHz) radio waves over great distances
has always been exploited in communications, and sometimes frequencies
lower than HF are used, as listeners to long-wave radio will know. During
World War II the UK air defence radar 'Chain Home', operating on
20-30 MHz, was occasionally troubled by 'nth-time-around' clutter created
206
OVER-THE-HORIZON RADAR 207
when the radio signal was scattered by the ionosphere and travelled unusually
long distances. Under these conditions the normal operating PRF of 25 Hz
was reduced to t 2.5 Hz (details in Neall). In other countries, similar
discoveries were made, and many radar engineers began thinking of turning
this unwanted propagation to advantage.
From the early experiments with long-range propagation, two kinds of
HF radar or 'over-the-horizon' (OTH) radar have been developed, known
as surface-wave (or groundwave) radar and skywave radar, making use of
the surface-wave and the ionospheric modes of propagation respectively.
Surface-wave systems were first operated in the early t 950s and effective
skywave systems a little later. The reason for the slow development of OTH
radar compared to more conventional systems is not a reflection on the
ability or the imagination of the engineers involved but rather the constraints
of the technology available at the time.
Surface-wave radar uses the surface-wave propagation mode to look
over the immediate horizon, and it may be used to survey ranges up to a
maximum of certainly no more than 400 km. It is most useful as a local area
defence system and as a method of collecting good-quality wave and tidal
information over a restricted area of ocean. Although bistatic systems have
been operated successfully, surface-wave radar is regarded as being a
predominantly monostatic technique with a relatively low capital cost.
Skywave radars, on the other hand, are almost always large, bistatic
and very expensive. These radars make use of the ionosphere to scatter radio
waves very long distances beyond the horizon, sometimes in forward scatter
mode to a receiver beyond the target. The minimum range is about 1000 km
and the maximum useful range is around 4000 km. Skywave radar is thus
more suited to the defence and remote ocean sensing needs of countries of
such continental proportions as the USA, the former USSR and Australia.
the transmitted signal (see Fig. 10.1). The vector sum of the large vertical
transmitted signal and a small induced horizontal field causes the resultant
wavefront to tilt over such that the Poynting vector (the direction of energy
flow) enters the sea at the Brewster angle t. One way to imagine this is to
picture the lower part of the wavefront having a slightly lower velocity due
to the water, dragging behind, and so bending the beam downwards. This
propagation mode has various names, but it is possibly best known as the
Norton surface wave.
The block diagram of a typical gw radar is shown in Fig. 10.2.
'Floodlight' illumination of a sector of sea is provided by a log-periodic
transmitting antenna, which has vertically radiating elements of varying
length to enable it to operate over a wide bandwidth. The receiving antenna
is a 100 m array of monopoles parallel to the coast, which are used to form
a number of beams simultaneously. If the system were installed in a confined
area such as on an oil-rig or ship, the transmitting antenna would almost
certainly have to be a monopole as well.
The transmitter is usually a wideband linear amplifier chosen from a
range of commercial HF communication equipment. Unlike pulse trans-
mitters, which can only deliver their peak power in short bursts, com-
munications transmitters are continuously rated, meaning that they can
deliver their specified power output continuously. The SNR of an HF radar
system can thus be improved by finding waveforms that extend the percentage
of time that the transmitter is on.
t The Brewster angle, named after a distinguished Scottish physicist of the last century. Sir
David Brewster, is best known in elementary optics. A beam of light striking a glass block at
a steep angle of incidence will enter and be refracted, whereas a beam striking at a very oblique
angle will be reflected from the surface. In between these two cases lies the Brewster angle, at
which the light runs along the interface between the glass and the air. In the case of gw radar,
the air/sea boundary acts in a similar way to the air/glass interface, and the radio wave travels
akmg the sea surface.
OVER-THE-HORIZON RADAR 209
Low-phase
noise synthesizer
Similar to a low-
frequency spectrum anlayser'
If the FFT has M points (usually a power of 2), we can relate the number
of pulses transmitted during the integration time tj to the number processed
by the receiver as
(PRF) x tj = nM [ ] ( 10.1 )
Range [km]
Figure 10.3 The additional two-way loss for propagation over a smooth surface: four-thirds
Earth, 40hm- 1 m-I conductivity. Frequency [MHz] is shown next to each curve.
Despite all the problems of HF gw radar, the rewards for being able to
stay in the safety and comfort of the shore, and yet monitor the sea and sky
beyond the horizon, are such that there remains considerable interest in this
form of radar.
Skywave (sw) radar makes use of scattering from the ionosphere to look
down on a 'footprint' well beyond the horizon. The main applications lie
in detecting ballistic missile launches and tracking military and civilian air
targets. Because wind direction is relatively easy to extract from sea clutter,
storm tracking is also possible. Most skywave radars also have a significant
capability for detecting surface targets and for sea sensing. Cruise missiles
are not easily detected because they have a small RCS when viewed at long
wavelengths. The possible exception to this is submarine-launched cruise
missiles, which leave the water vertically and for a few moments may present
a large RCS before they settle into horizontal flight.
Unlike gw radar, skywave systems do not need to be near the sea and
are perhaps best located inland, where they are safe from storms, salt spray
and enemy attack. Skywave radars also tend to use up a lot of ground space,
and it is often easier and cheaper to find suitable sites inland, especially since
the receiver needs a radio-quiet location. The receivers and transmitters are
almost always separated, often by as much as 100 km, thus permitting the
use of FMCW modulation to maximize the mean power output of the
transmitters.
A typical arrangement for a sw radar is shown in Fig. 10.4. A number
of continuously rated high-power transmitters amplify an FMCW waveform
and drive a total power of about 500 k W into a 200 m long array. The type
of transmitting antenna is chosen to provide a good impedance match to
the power amplifiers over a wide bandwidth and to have a beamwidth
somewhat wider than the receiving antenna. There are two reasons for using
a relatively wide transmitting beam: first, the dwell time may be increased
if several narrow receive beams are used simultaneously to survey the area
illuminated by the transmitter array (Fig. 10.5); secondly, for a given antenna
gain product, the clutter cell size is smaller when the two antenna gains are
unequal.
Continuous scanning is not usually possible with sw radar because a
finite amount of time is needed to change the beamformer of the transmitting
antenna to a new position. It may also be necessary to retune to a new
frequency to illuminate a different range, in the same way that HF
communications engineers have to. For surveillance of a wide area, a step
scan technique is used, shown in Fig. 10.5, in which the radar illuminates
each sector for several seconds before moving on to the next. As is usual
scatter
receiving
station
Frequency FMCW
management waveform
system generation
"
"
"
"
"
'I
"
N "
~
- Figure 10.4 Typical skywave radar installation.
214 UNDERSTANDING RADAR SYSTEMS
2
Sectors illuminated
by transmitting ,
•
------------ I
antenna \ ---------- ____ l ~
, ------------- I
t :::::::::::..-___---- J _ Sectors.s~rveyed
• _____----- __ ----, by receIVIng
-
, ------=-::.-.:.:: - antenna
Figure 10.S Step scanning is used to survey a wide area. Each sector is illuminated for a few
seconds and surveyed by four narrow receiving beams, before moving on to the next sector.
with surveillance, the objective is to scan a wide area and return to the
starting sector before a target can pass through it undetected.
The transmitting array elements are typically vertically polarized log-
periodic, although the North American CONUS-B system uses short, fat
dipoles canted 45° off-vertical ('fattening' a dipole increases its bandwidth).
The elevation pattern of the array depends on the impedance of the ground,
which in turn can be affected by the weather, so usually the ground is levelled
and some form of conducting wire mesh is laid down to stabilize the
impedance. The improved ground conductivity also helps to predict and
control the antenna sidelobes.
The receiving antenna is usually a long (greater than 1 km) array of
monopoles, sometimes with a backscreen to reduce the backlobe and
occasionally with the monopoles fattened to improve the match at low
frequencies. Often, antenna losses at low frequency are not important because
the system is externally noise-limited, meaning that the external noise
(galactic, atmospheric, man-made) arriving with the signal is greater than
the internal noise of the receiver. In these circumstances, some losses at the
antenna are permissible because noise and signal are attenuated equally and
the SNR remains unaffected. But the losses must not be allowed to become too
severe-in HF engineering it is regarded as something of a crime to permit
losses so large that a system becomes internally noise-limited.
The receiving array is connected to a beamformer. This may be an
analogue device, but more often these days this function is performed as a
OVER-THE-HORIZON RADAR 215
digital process using one receiver plus A/D converter for each antenna
element, or for each sub-array group of elements. Analogue beamformers
steer the beam by introducing a frequency-independent time delay (see
Chapter 15); digital beamforming usually involves a phase delay that is
equivalent to the time delay at one specific frequency f (the operating
frequency). For this reason, phase delay beam steering has a restricted
bandwidth B, given roughly by
B / f [%] ~ ~8 [ degrees] (10.2)
where ~8 = antenna beamwidth. This bandwidth restriction also implies a
limit to the range resolution ~R and a similar working rule of thumb is given
by
~R ~ D sin 8 Em] (10.3 )
where D = length of the antenna array Em] and 8 = beam angle off-bore-
sigh t [degrees].
After the beamforming and receivers, the signal and data processing is
based extensively on the mapping of the doppler spectrum into range/azimuth
resolution cells. These cells are searched using an adaptive thresholding
process, and any targets detected are tracked making use of the doppler
measurements to compensate for the relatively poor range and azimuth
information. The azimuth resolution is poor because even a 1 beam diverges
0
Resolution
Radar Power (typical)
F2 region
E region
L-------f~o~E~~fu~F~1--------~fo~F~2~----+
layer persists during the night with the ionization falling to around
5 x 10 10 [electrons m- 3 ] [foF2 = 2 MHz).
The bottom of the ionosphere is not smooth, but has a roughness
comparable with the scale sizes of the hills and valleys on the earth's surface,
with the added complication that the roughness is moving and changing
with the background neutral wind and electric fields. All three ionospheric
layers are subject to various disturbances, irregularities and anomalies, and
the electron densities may vary with time of day, season, location and solar
activity. The ionosphere is thus an uncertain and ever-changing medium of
propagation, and there may be many paths by which a signal can pass from
the transmitter to the target and back to the receiver-this can cause a single
target to appear at many apparent ranges (see Fig. 10.7). There is also a
limited band of frequencies that can be used to illuminate a target at any
given range and, as with HF communications, the band required may be
congested with radio traffic and high background noise levels.
F2 region
Fl region
iii' 50
x
:2.
c.:::
zVl x x
0
."
'"
>
.;;
-50
X
x x
u
'"
c.:::
4000
Apparent range [km]
Figure 10.7 Radio signals may follow any combination of E and F region outward and return
paths to give several apparent ranges and values of SNR for a signal target.
OVER-THE-HORIZON RADAR 219
The answer to the problems of using the ionosphere for OTH radar lies
in a process known as frequency management. Extensive mathematical
modelling and continuous observations of the ionosphere are used to select
the best frequency band and then wide band 'look-ahead' or channel occupancy
monitoring receivers are used to find a channel free of interference. ,
Over the years, quite realistic mathematical models of the ionosphere
have been constructed (as computer software) to include both local data-
bases of observations built up over several solar cycles and maps of the
mean ionosphere produced by the CCIR (International Radio Consultative
Committee ). Statistically, these models provide a good prediction of expected
conditions, but on any given day the vagaries of the ionosphere are such
that the prediction can be substantially in error. These errors can be
minimized by ionospheric sounding, which measures the current state of the
ionosphere as an aid to selecting the most appropriate model. Occasionally,
part of the radar system itself is used for oblique sounding, but it is more
usual to use a vertical sounder, or ionosonde, a small pulsed radar system
that transmits upwards and sweeps in frequency to locate the critical
frequencies of the E and F layers.
The most common form of display for sounding data is an ionogram,
a plot of virtual height (not always equal to the true height because of the
slowing of radio waves in the ionosphere) against frequency, as shown in
Fig. 10.8. After some calculation, Fig. 10.8 can be replotted as true height
against electron density to produce a form suitable for ray tracing programs
to predict where signals transmitted on any given frequency will end up.
This is the information needed by sw radars to choose the best operating
frequency.
There are practical methods of ensuring that the frequency management
program is working as predicted. The radar operator can ensure that easily
Frequency [MHz)
recognized targets (cities, coastlines, islands) are detected and appear in the
correct locations. Also known targets and transponding devices can be placed
within the coverage area to give confidence in the terrain illumination and
some feedback on the ionospheric absorption.
Ionospheric modelling, frequency management and calibration form a
major part of skywave radar operation and contribute to the expense of
these large surveillance radars.
SNR = PtGtGr(J'),2{Js/it/ir
[ ] (10.4 )
(4n) 3 R~ R; FakTo
where Pt = mean transmitted power [W]; GtGr = antenna gains relative to
an isotropic radiator in free space [ j; (J' = target cross-section, not
free-space but as measured, i.e. including any surface reflection effects [m2];
A. = radar wavelength Em]; ti = coherent integration time [s]; Is = system
loss factor [ ]; lit, lir = additional path losses on outward and return paths
due to D and E region absorption [ ]; R t , Rr = outward and return
distances to target Em]; and FakTo = apparent external noise level [W Hz-I],
with Fa the effective antenna noise factor, usually expressed in dB above kTo,
i.e. above - 204 dB W Hz - 1. The bandwidth is assumed to be 1/ t i •
Except during auroral absorption the loss factors lit and lir may amount
to only a few decibels during the day and are near zero at night (details of
how to calculate these losses may be found in reference 4 ). System losses are
more severe and are due to a variety of causes related to the antenna design
and signal processing. Typical system losses are listed in Table 10.2 and,
while they may vary from one system to another, they seldom fall below 10 dB.
The effective antenna noise factor Fa (the result of external noise)
increases with increasing wavelength, and over the range 3-15 MHz a
realistic value to choose for back-of-the-envelope performance estimates is
Fa ~ 60 - 2fMHz [dB] (10.5 )
so that the noise level in a bandwidth B [dB Hz] is given by
Next = 60 - 4fMHz - 204 +B [dBW] (10.6 )
More detailed formulae, representing different noise environments in the
USA, are to be found in Skolnik 5 . The classic work in this field is the CCIR
Report 322 6 , which, although published in 1963, remains a very useful
database of global atmospheric noise levels.
OVER-THE-HORIZON RADAR 221
The ground-wave variant of the radar equation differs from the sw version
because the presence of the conducting ground modifies the antenna gains
and creates the ground-wave effect (see Hall and Barclay2 and Shearman in
Scanlan 7). A frequently used version of the gw radar equation is
SNR = ~G;G;a'tJs161/
[ ] (10.7)
(4,.)3R4F.kTo
Besides the propagation difficulties, there are four other problem areas
encountered when operating both gw and sw HF radar. First, there are
problems associated with licensing powerful HF transmitters and with getting
the frequency allocations clear of other traffic that are wide enough to give
good range resolution. These difficulties should not be underestimated.
Secondly, there are often siting difficulties. It can be difficult to find
sufficiently large and level radar sites that meet the requirements of being
radio-quiet at the receiving end and of presenting no sensitive environmental
issues at the transmitter location.
Thirdly, both gw and sw radar experience ionospheric clutter (unwanted
backscatter from the moving ionosphere) and meteor clutter, which can at
times swamp the presence of wanted echoes. The extent of the meteor echo
problem can be evaluated using the equations given in Chapter 9. Skywave
radars can also suffer significant polarization losses and focusing/defocusing
due to ionospheric effects.
Lastly, military HF radars are susceptible to deliberate jamming and
only modest powers are required to disable them, even from relatively long
ranges. However, OTH radars are usually classed as early warning devices,
rather than as accurate tracking radars, and so they probably fulfil their
warning role to some extent if they report that electronic warfare has been
initiated.
10.7 SUMMARY
There are two forms ofOTH radar, surface-wave and skywave. Surface-wave
systems are relatively inexpensive and have found applications in sea sensing
(see next chapter), for the defence oflocalized areas against low-flying missiles
and to some extent for monitoring ship traffic.
Skywave radars are used to monitor very large areas of land and sea to
search for air targets, ballistic missiles during launch phase and some types
of surface target. They also have remote sensing capabilities, especially storm
tracking and ocean wave monitoring. These radars are large, powerful,
expensive and require sophisticated frequency management systems in order
to operate via the ever-changing ionosphere.
Key equations
• For pre-summation and fast Fourier transformation:
(PRF) X tj = nM []
224 UNDERSTANDING RADAR SYSTEMS
10.8 REFERENCES
I. CH-the first operational radar, B. T. Neal, GEe J. Res., 3(2), 73 - 83, 1986.
2. Radiowave Propagation, Ed. M. P. M. Hall and L. W. Barclay, Peter Peregrinus for the
lEE, Stevenage, Herts, 1989. [There are chapters on both gw and sw propagation.]
3. Jane's Radar and Electronic Warfare Systems 1989-90, Jane's Information Group, London,
1989.
4. Over-the-Horizon Radar, A. A. Kolosov et al., Artech House, MA, 1987. [This book attempts
to layout the foundations of OTH radar engineering, but the coverage is patchy and biased
towards single-hop skywave backscatter systems. The WARF radar is extensively cited.]
5. Radar Handbook, Ed. M. I. Skolnik, McGraw-Hili, New York, 1990. [Chapter 24 is dedicated
to HF OTH radar.]
6. World Distribution and Characteristics ofAtmospheric Radio Noise, CCIR Report 322, 1963.
7. Modern Radar Techniques, Ed. M. J. B. Scanlan, Collins, Glasgow, 1987. [Chapter 5 is a
useful review of skywave propagation, clutter problems and HF remote sensing by Professor
E. D. R. Shearman.]
8. Real-time sea-state surveillance with skywave radar, T. M. Georges, J. W. Maresca, J. P.
Riley and C. T. Carlson, IEEE J. pcean Eng., OE-8(2), 97-103,1983.
10.10 PROBLEMS
10.1 SRI International have used the Wide Aperture Research Facility (WARF) in California
to explore the benefits of high range and azimuth resolution for skywave OTH radars. A unique
feature of WARF is a 2.5 km long receiving aperture. The system operates over the range
6-30 MHz but, for all the problems below, assume a typical operating frequency of 15 MHz:
(a) What is the azimuth beamwidth?
(b) The elevation beam width is 36° between half-power points. What is the antenna gain?
(c)Two 10 kW transmitters feed an antenna of gain 20 dB. What is the total power x aperture
product p'G,G,?
(d) What is the figure of merit when an integration time of 12.8 s is used?
10.2 (a) If the transmitter in problem 10.1 is of the FMCW type and sweeps from 15.000 to
15.050 MHz, what is the area of water surveyed by the radar at a range of 1500 km?
(b) Using the definition in Eq. (10.8) of the reflectivity of sea water as -23 dB m 2 m- 2 ,
what would be the RCS of the sea echo?
10.3 What noise level would you expect to find in the problem above?
10.4 What SNR should be received for the sea echo in the problems above? Assume the
propagation term (fh)') is 10 dB worse than (4nR)2 /)..
10.5 An aircraft is detected in the same resolution cell as the sea echo in the problem above.
If the tracker reveals the aircraft to be approaching the radar with a doppler shift of 15 Hz and
a SNR of 23 dB, what can you deduce about the type of aircraft?
CHAPTER
ELEVEN
RADAR REMOTE SENSING
The way radio waves scatter from the earth's surface can be used to
investigate large-scale features of land and sea.
Random, noise-like processes are often interesting, but the surface of the sea
has a particular fascination. People stare at it for hours. For a radar engineer,
the challenge is first to find out how radio waves are scattered from the sea
and then to turn the problem round and ask, 'How can we determine what
is happening at sea from our radar observations?' In some ways, this type
of inverse problem is like being given an answer and having to find the
question.
What is an ocean wave? We must answer this question before we can
make any progress with the problem. Wave motion is carried by particles
of water exhibiting circular motion as the wave travels past. This is easily
demonstrated by floating a cork, or small piece of wood, on the sea and
watching it move as a wave travels by. At the surface of the sea these circular
motions can have quite large amplitudes, but they quickly die away with
226
RADAR REMOTE SENSING 227
Wave propagation
•
Figure 11.1 The trochoidal shape of sea waves generated by the circular motion of water.
depth (see Fig. 11.1). Submarines soon avoid the effects of wave motion
when they submerge.
Figure 11.1 shows why the shape of the wave itself is not a sinusoid but
is a rather flat-bottomed/pointed-crest type of shape known as a trochoid.
The height of a sea wave is small compared to its length; a typical wave
steepness is 1/18 for open ocean sites, with 1/10 rarely exceeded.
Waves are generated by the wind either locally or by distant storms.
These latter waves, known as swell, have long wavelengths, which enable
them to travel great distances with low attenuation. Locally driven waves
are more sophisticated. The wind blows over the water and begins to form
short-wavelength waves, but as these build up, and processes become
non-linear, the energy is transferred into longer waves with larger amplitudes.
For a constant wind speed, an equilibrium is reached with the largest waves
having a speed close to that of the wind. If the wind strengthens, then longer
and larger waves will be generated, giving the type of spectrum shown in
Fig. 11.2.
20 knots=sea state 4
\0 knots
0.2
Frequency [Hz)
Figure 11.2 A typical waveheight spectrum. This is a non·directional spectrum because it gives
no information about wave direction. In fact, the largest waves tend to be aligned with the
driving wind force but the smaller waves are omnidirectional, giving rise to the complex 'choppy'
look of the sea surface.
228 UNDERSTANDING RADAR SYSTEMS
- - - - - - - - ~----- -- - - - -------
--- -------- --~----- - -- ~ -------=-----: -----~--------- --- -------------=-
Figure 11.3 When the sea wavelength equals half the radar wavelength, the backscatter from
one crest has a path length shorter by a whole radio wavelength than the backscatter from the
next crest.
COMMENT Since most radar systems use range cells many times larger
than the radar wavelength, this enhancement due to Bragg scattering
becomes the dominant mechanism irrespective of the radar frequency
used.
RADAR REMOTE SENSING 229
/
/
/ e
~-------------
L"
--
--- -- -- --"'------ ---------- --=---::::- -- -- ----- -- -
-
Figure 11.4 Bragg resonance occurs when the path difference between the echo from two wave
crests is a multiple of }./2.
We now work out the doppler shift of the Bragg resonant sea waves, as
observed by the radar. To do this, three pieces of information are needed.
The first of these is the Bragg formula, which should be slightly more general
than the case shown in Fig. 11.3 because the radio wave could be incident
on the sea surface at an oblique angle. The full case is shown in Fig. 11.4
and the following equation:
Ln = nA/(2 sin 0) [m] (ILl)
and for groundwave radar sin 0 = 1. Note the possibility that signals
back scattered from the sea waves of wavelength 2 x )./2, 3 x )./2, etc., will
also add coherently at the receiver.
The second piece of information needed is the dispersion relation for sea
waves, which relates the phase velocity Vn to the sea wavelength Ln:
v" = J (g L n/ 2n ) ( 11.2)
where g is the acceleration due to gravity.
Finally we need the doppler formula of Eq. ( 1.20), derived in Chapter 1:
in = ±(2v" sin 0)/). [Hz]
Putting these together (an exercise left to the reader!) gives
-10
a:l
/Yz X Bragg
~
Receding Bragg line /21 x Bragg
t -20
~
oQ.
<l)
.~ -30
.,
(;j
0::
-40
o 2
Doppler/Bragg frequency
The effect of a tidal current in the sea is to shift the whole of the doppler
spectrum by an amount equal to the radial component of the drift. In practice,
current measurements are often made by measuring the displacement of just
the two Bragg lines. A single radar can only measure the radial component
of the current and it is common practice to use two separate radars, with
overlapping beams, in order to derive the full current vector information.
There are now several commercial HF radar systems for measuring tidal
currents. Even a short-range system, measuring only coastal currents, can
have many applications because of the need to understand more about such
problems as coastal erosion, sandbank movements, sewerage distribution,
warm effluent distribution (from power stations), docking large container
ships and stress on dykes during tidal surges.
The measurement of the Bragg line displacement represents a weighted
average of the current over the top layers of the ocean. The radio wave itself
does not propagate far into the water (the attenuation with depth being
about 60 dB per metre at 3 MHz and 189 dB per metre at 30 MHz). However,
the orbital motion of the water particles, shown in Fig. 11.1, means that the
radar is sensitive to subsurface currents at depths up to 4 m at 3 MHz and
0.4 m at 30 MHz.
Surface currents measured remotely by HF radar agree well with in situ
measurements made by current meters and drift measurements and are
accepted as being reliable. HF radar also has the advantages of being able
to survey a wide area of sea simultaneously and of being unaffected by bad
weather and rough seas. It is sometimes possible to gain further insight into
small-scale surface current structures by examining any broadening of the
Bragg lines or by using several frequencies simultaneously.
Bistatic Operation
Most HF sea sensing systems now use at least two radars to observe a single
patch of water in order to increase the amount of information on the
two-dimensional sea wave spectrum. If two radars are operated together,
further information can be derived by also operating them bistatically such
that transmissions by one are received by the other. This arrangement gains
a third view of the sea spectrum along a line that bisects the angle between
the two radar beams. In future, this may help to increase the accuracy with
which wave information can be obtained from HF radar observations.
Sea Ice
Detecting the presence and movement of sea ice is possible using HF radar,
and to some extent the thickness of ice may also be estimated. The penetration
of radio waves into sea ice is somewhat variable because ice is a complicated
substance whose dielectric properties depend on such things as the tempera-
ture, the age of the ice and the amount of brine trapped within it.
Skywave mapping of the Greenland ice cap shows that thin ice has a
reflectivity similar to that of sea water, and the contour of reflectivity 10 dB
lower corresponds to an ice thickness of 1000 m. Groundwave observations
show pack ice reflectivities between 2 and 10 dB greater than the reflectivity
of the sea, and icebergs appear to be 12 or 13 dB greater.
Ice is a 'hard' target, and drift speeds may be inferred directly from the
doppler shift-just as for a ship or an aircraft-although long integration
times may be necessary to detect slow ice movements. An alternative
approach is to plant transponders on the ice to increase the accuracy of the
velocity measurements by imposing a small calibrated frequency shift on the
echo. This frequency shift makes it easier to distinguish the ice movement
from echoes due to 'stationary' targets such as land, stationary ice,
second-order sea clutter and transmitter-receiver breakthrough.
Worked example If the doppler power spectrum shown in Fig. 11.5 were
obtained from an HF groundwave radar operating at ! MHz and an
echo from a transponder antenna were just visible as a spike at the centre
of the spectrum, what would be the improvement in signal/sea clutter
ratio if the antenna impedance were modulated such that half the signal
appeared at a doppler shift of 0.85 Hz?
SOLUTION Substituting 8 MHz into Eq. (11.3) shows that the Bragg
lines in Fig. 11.5 must have a doppler shift of 11 = 0.29 Hz (sin (J = 1
234 UNDERSTANDING RADAR SYSTEMS
Ship Tracking
A necessary part of wave sensing is the removal of 'ship clutter' appearing
as spikes in the doppler spectrum and which may be confused with
second-order sea echo features. Conversely, when tracking ships, it is necessary
to remove sea clutter. There seems to be a good case for developing integrated
systems with two-stage data processing. First, recorded data would be
scanned for the presence of ship echoes, which would then be tracked through
the recordings, removed and the information transferred to the appropriate
users. Secondly, the ocean wave and current information could then be
derived from the cleaned-up data, and distributed to interested parties.
Besides the obvious military interest in ship detection and tracking, there
are also civil applications for the control of shipping. This already takes
place in small, densely populated shipping areas using microwave radar
systems, but there is no long-range control of shipping equivalent to air
traffic surveillance.
With the wide range of applications for HF radar remote sensing and
the possibility of combining many of these functions into a single system, it
is likely that this will remain an active and interesting field of radar
research for many years.
2.00
1.50
1.00
0.50
Figure 11.6 The approximate gaussian PDF obtained by averaging measurements from
exponential clutter; the shaded areas must have combined area less than 5 per cent to give 95
per cent chance of estimating the mean within a tenth of its value.
RADAR REMOTE SENSING 237
PDF to be of area 0.025. Using Table 4.1 and the associated figure, we
can see that this occurs when (f)( t) = 0.475, so that the distance from
the mean is 1.96 standard deviations. Hence JM = 19.6, or we need
384 independent samples. Such large sample sizes are normal if very
precise measurements are needed in exponential fading.
,/
\ /
\ /
/
225km
~~~~rometer ~\
/
/
/ /
/
antennas \ /
/ \ /
/ 500km
/ \ /
// \ /
/ \ /
/
/ \
// /
k----"'-----'( Aft beam
2}/
~'1i'/
x,~ /
0/
Figure 11.7 The geometry of the space-borne scatterometer carried on the ERS-l satellite.
approximately planar within the radar footprint. This is the case over the
ocean and the earth's great ice shelves. One such device was carried on the
Seasat satellite. It provided remarkable images showing that the ocean surface
has a topography reflecting the trenches and submerged mountain ranges
in the deep oceans. Large-scale currents such as the Gulf Stream can also
be extracted from the data because they give rise to deviations from the
mean height.
An altimeter operates by emitting a pulse vertically downwards. A
pulse-limited altimeter uses a very short pulse, and the sequence of events
that follow is as outlined below:
1. The leading edge of the pulse strikes the surface, and a return signal
begins to propagate back to the receiver.
2. The illuminated area grows like an expanding disc, until the trailing edge
of the pulse reaches the ground. As the illuminated area grows rapidly
in size, so does the returned power.
RADAR REMOTE SENSING 239
This sequence and the associated returned power is illustrated in Fig. 11.8.
The ERS-l satellite carries an altimeter that operates at 13.8 GHz, using
an antenna of 1.2 m diameter. In order to avoid the generation of very short
pulses and the associated rapid processing, a slightly different operating
strategy is used. A chirped pulse is transmitted, and the analysis of timing
and shape is carried out in the frequency domain. The essential principle is
unchanged. Because the altimeter operates on a single frequency, the
correction for atmospheric delay (see Chapter 8) must make use of iono-
spheric models or ancillary data.
The returned waveform in fact yields more information than just the
time delay to the surface. Over the ocean, which is anything but a flat surface
Radius of annulus
Trailing edge grows, but
reaches surface - Illuminated area is constant.
Leading edge Illuminated area disc has region becomes until edge
strikes surface grows maximum area an annulus of beam reached
Figure 11.8 The sequence of events that occurs for a single pulse of a pulse-limited altimeter
system, and the expected form of the return at the satellite.
240 UNDERSTANDING RADAR SYSTEMS
on the length scales the altimeter is measuring, the midpoint of the slope of
the leading edge of the return corresponds to the mean distance to the sea
surface. This is effectively an indication of the geoid, on which the earth's
gravitational potential is constant. The slope of the leading edge of the return
is determined by the spread of surface heights in the altimeter footprint. This
gives information on the variance of surface height, i.e. waveheight. The total
returned power depends on the small-scale roughness of the ocean surface,
which depends on the wind speed; this is the dependence that a scatterometer
exploits.
Altimetry can also be used over the ice shelves. This aspect of altimeter
operation is less well understood, because of the possibility of penetration
of the surface by the radar wave and because we do not have a simple
description of the surface, as we do for ocean waves. Nonetheless, this is an
active area of research and experiment usirtg the ERS-l altimeter, because
the dynamics of the ice shelves are a critical indicator of the effect of global
warming. They also play an important part in the amount or energy absorbed
by the earth from the sun, and in earth-atmosphere interactions.
Perhaps the most sophisticated radar technique (certainly the most tech-
nically demanding) used in remote sensing is SAR. The major attractions of
SAR are that it can provide high-resolution images of extensive areas of the
earth's surface from a platform operating at long ranges, irrespective of
weather conditions or darkness. The resistance to weather conditions derives
from the use of wavelengths of the order of centimetres, with the X-band
(3 cm), C-band (6 cm) and L-band (24 cm) being favoured. Some systems
use shorter wavelengths (down into the millimetre bands); these will be
adversely affected by precipitation and cloud, for the reasons discussed in
Chapter 8. P-band (68 cm) SAR is also in operation. A few reservations are
needed here. For air-borne SAR operating at longer ranges (see Fig. 11.9),
the propagation path may have to traverse extended regions of rain in the
troposphere. The effects discussed in Chapter 8 can then affect performance.
For space-borne SAR, the ray paths through the troposphere are much
shorter (see Fig. 11.9), and only the most extreme tropical rainfall could
have any effect, and even then only at the shorter wavelengths. However,
depending on the orbit height, the path lengths through the ionosphere may
be long. In the post-sunset equatorial regions and in the auroral zones,
ionospheric irregularities can destroy the phase coherence essential to SAR
performance. Examples of the consequences for the antenna pattern of a
SAR were shown in Fig. 8.11.
This all-weather capability makes SAR a most attractive tool for
environmental monitoring in regions affected by clouds or darkness and for
RADAR REMOTE SENSING 241
_----t ~----------_______ _
-- ---- Satellite altitude
(up to several
thousand kilometres)
--.------
-- --
----------, , Ionospheric
peak electron
density (- 300 km)
Aircraft
"-
"-
"-
" '- Tropopause
(- 15 km)
Figure 11.9 A comparison of the propagation paths of air-borne and satellite-borne SARs.
The principle by which SAR obtains high spatial resolution was first
pointed out by Carl Wiley of Goodyear in 1957, and is in fact very
straightforward. Consider a moving platform (aircraft or satellite) that
carries a pulsed radar pointing sideways to its motion. At a given range, the
antenna illuminates a strip of scatterers; as it moves forwards, new scatterers
enter the beam, and others leave it (see Fig. 11.10). For a single scatterer,
on each pulse of the radar, the two-way propagation to and from the scatterer
causes the phase to change by
[rad] (11.4 )
Here S, is the distance between the radar and the scatterer when the lth pulse
is emitted.
Buried in the sequence of returns during which the scatterer is illuminated
by the antenna is this phase history, which is unique to a scatterer in that
position. By storing the sequence of returns and correcting for this particular
phase history (by subtracting V't from the phase of the lth pulse), the returns
from the scatterer can all be brought into phase. Adding all the corrected
,,
,,
Scatterer entering
,, " the beam
Scatterer leaving
the beam
Figure 11.10 The area illuminated by the SAR antenna.
RADAR REMOTE SENSING 243
pulses together will cause the returns from the scatterer to add up con-
structively (see Fig. 11.11 ). For scatterers with different phase histories, this
process will be destructive, and they will contribute far less to the summation
than the scatterer to which the phase correction is 'tuned'. Scatterers that
are close together in the along-track direction will have similar phase
histories, and they will have comparable weightings in the summation. The
separation that causes one of them to suffer significant destructive interference
when the processing is tuned to the position of the other is a measure of the
system resolution. From Fig. 11.11 we can see immediately that the SAR
simply operates like a big lens or antenna array. The only real difference is
that the individual rays through the lens are gathered sequentially rather
than simultaneously. In doing this, a large-aperture antenna is synthesized.
Large apertures lead to good angular resolution, and this improvement
Sum coherently
Correction of
phase
Pulses emitted
d<Vk = phaseshift
along two-way
propagation path
Synthetic antenna
gain pattern
Figure 11.11 The SAR principle from the viewpoint of correction of a sequence of phase delays.
The beam pattern of the synthesized antenna is indicated at the bottom of the diagram.
244 UNDERSTANDING RADAR SYSTEMS
(a) (b)
Flight track
~
---+
D
[rad] (11.8 )
The angular beam width of the SAR is inversely proportional to range. But
the spatial resolution on the ground will be the product of the angular
resolution and the range. Hence the along-track resolution of the SAR is
ra=R{3s=d/2 Em] (11.9)
This is a key equation. It tells us that the ground resolution of the SAR in
the along-track direction is independent of the range and also of the
wavelength. It also illustrates the curious property that, in order to obtain
better resolution we should shorten the antenna, in complete contrast to
what we expect for real antennas. This property arises because shorter real
antennas give rise to longer synthetic antennas (see Eq. ( 11.7». In fact, other
operating characteristics, such as the width of the usable swath, impose
constraints on the minimum length of the antenna 2 (see problem 11.5) and
hence on the along-track resolution. Equations (11.8) and ( 11.9) should be
compared with what happens in a conventional radar for which {3 is given by
Eq. (11.6), so that the alo~g-track resolution at range R is R{3 = RA/d.
Worked example Compare the azimuthal resolution of a conventional
radar antenna and a focused SAR for a radar frequency of 3 GHz and
an antenna size of 10 m. Calculate the results for ranges of 10, 100 and
1000 km.
SOLUTION At 3 GHz, the wavelength is 0.1 m. For the focused SAR, the
resolution is independent of range at d/2 = 5 m. The resolution of an
ordinary antenna varies from 100 mat 10 km range to 10 km at a range
of 1000 km.
The discussion above only relates to the along-track direction. Cross-
track resolution is obtained by transmitting a chirped pulse and pulse
compression. In fact, the along-track processing can also be thought of as
linear FM with pulse compression. To see this, consider how the range to
a stationary scatterer varies as it passes through the beam. Using Fig. 11.12b
we can write
(11.10)
t If you are familiar with antenna theory, you will know that Eq. (11.6) is valid in the lill"
field of the antenna. For SAR, scatterers are in the far field of the real antenna but in the near
field of the synthetic antenna. That causes the quadratic phase variation discussed below.
246 UNDERSTANDING RADAR SYSTEMS
,1,( ) _
'I'
2nV2 2
t - ---t [rad] (11.13)
A.R
(The minus sign occurs because we must subtract the phase delay to get the
phase; the absolute phase reference is unimportant, and we have set it to
0.) Frequency is the time derivative of phase, so this is a linear FM signal;
the instantaneous frequency is
4nV2
w(t) = - - - t (11.14 )
AR
We arrived at exactly the same result from a doppler frequency point of view
in Chapter 6 (see Eq. (6.35».
From Eq. (11.9), we can see that SARs should have spatial resolutions
of the order of metres using practical antennas. (For example, the C-band
antenna on ERS-l is of length 10 m, giving a theoretical resolution of 5 m.
When a more precise defiinition of resolution is used, and the antenna and
processing weighting is taken into account, the quoted resolution is 6 m.)
What price is being paid for this resolution? There are two principal costs.
The first of these is that each pixel represents only a single sample from
what is, for many types of surface, an extended clutter-type target giving
exponential fading. This is the speckle phenomenon, which causes major
problems in SAR image interpretation. Many systems choose to sacrifice the
highest resolution possible from the system in order to reduce the effect of
speckle. They do this by averaging the detected powers from several
neighbouring pixels (either along- or cross-track, or both). This causes the
resolution to be degraded; averaging M pixels in either of the directions
degrades the resolution by the same factor in that direction. The gain in
accuracy (and reduced variability) of the measured RCS comes about because
the pixel distribution changes from an exponential to a chi-squared distri-
bution with 2M degrees of freedom, where M is the total number of pixels
being averaged for each measurement. As we already remarked when
discussing scatterometry, this distribution tends to normality when M
RADAR REMOTE SENSING 247
becomes larger; the mean Jl. is preserved, and the standard deviation becomes
Jl.IJM. This process of averaging powers (or, in some cases, amplitudes) is
known as multi-looking. Note that the averaging of pixels can be built into
the processing; the formation of the full-resolution pixels may never occur.
This is, in fact, what a scatterometer does; a scatterometer and a SAR are
very closely related instruments. (On ERS-l they are integrated in what is
called the Active Microwave Instrument; while sharing electronics, they use
different antennas because of the different operational requirements of the
scatterometer. )
The second cost is hidden in Eq. (l1.5). To obtain the best possible
resolution at range R, the SAR must save and process all the returns in the
synthetic aperture of length D. The number of these returns is determined
by the PRF. This is, in turn, determined by the Shannon-Whittaker sampling
theorem: the sample rate must be at least twice the bandwidth (in Hz). From
Eq. (l1.l4), we can see that the frequency sweep due to a scatterer is
determined by the time it is illuminated by the real antenna, i.e. the time
during which the platform travels a distance D. This is simply DIV. Hence
the total frequency sweep is given by
[Hz] (11.15)
Since the time between samples is I/PRF, this is equivalent to the condition
that the distance moved by the platform between samples must not exceed
d12, i.e. the radar must emit at least two pulses in the time the platform
moves a distance equal to the real antenna length.
This means that, to carry out the processing at range R, we must save
the returns from 2Dld pulses for each pixel we generate, i.e. 2R).ld 2 pulses
are needed. Everyone of these pulses must be phase-corrected. This implies
a complex multiplication (four real multiplications) for each pulse, using the
I and Q channels of the returned signal. Therefore, each image pixel at range
R requires 8R)'; d 2 real multiplications. Since new samples are produced at
the rate 2 V / d per second at the minimum sampling rate, a real-time processor
would require 16RV).ld 3 real multiplications per second at range R. The total
number of operations must also take into account the number of range gates
MR' If we ignore the variation with range in order to get an order-of-
magnitude estimate of the task, we arrive at the need for MRR V ).1 d3
multiplications per second for real-time imaging, even without the processing
corrections that are normal in most modern processors. Note the dependence
248 UNDERSTANDING RADAR SYSTEMS
The discussion above has assumed that the processing in the cross-track
direction does not add significantly to the processing cost. This is true for
many air-borne systems, which use surface acoustic wave (SAW) devices to
perform the compression of the chirped pulse on reception. For satellite
systems and the modern generation of air-borne systems, the whole process
is carried out digitally. This increases the estimate of the processing required
by three orders of magnitude or more. As a result, the processing requirement
derived in the worked example is modest by the standards of current systems.
(As an example, the data rate of the ERS-I SAR is 105 x 106 bits per second.
This amount of data cannot be stored on board, and the SAR can only
operate if it is within line of sight of a receiving station. All the processing
is carried out on the ground.) Consequently, most SAR processors work
several orders of magnitude slower than real time, and for satellite SAR
real-time processing is considered very much state of the art. Even most
air-borne systems do not operate in real time. This occurs because high image
quality relies on a number of corrections being applied in the processing.
Uncorrected images are often produced as a quick-look product, with a
precision product being produced later (and slower). In fact, real-time
processing was normal in most early systems, using optical processing
elements (see Kovaly3 for a survey of the early development of SAR ). However,
these do not offer the flexibility and control of image quality possible with
digital methods, and almost all modern systems use digital technology.
Despite these drawbacks, SAR offers our only possibility of reliable high
resolution monitoring of many parts of the earth's surface. This has caused
major efforts to understand the interaction of the transmitted waves with
the surface, which gives rise to the observed backscatter. Nowhere is this
more true than over the ocean. The nature of the imaging mechanisms by
which SAR reveals the presence of ocean waves has been a subject of heated
debate ever since the L-band SAR carried by Seasat revealed the immense
RADAR REMOTE SENSING 249
Considerable
backscatter to
satellite
I
To satellite
Direction of motion
of long wave
c= :> I
To satellite
I
To satellite
Positive
doppler shift
imposed
Figure 11.15 Velocity modulation of small waves by the circulation of water in larger waves.
crop types. It has been realized that effective use of SAR for this application
is likely to require multifrequency and/multipolarization systems. The ability
to gather and relate measurements gathered at different times also seems
essential. Most of the work in this field has been carried out using air-borne
radars. The use of SAR in forestry has also attracted much attention. This
has been accelerated by the focusing of the world's attention on the
importance of the great forests for the long-term habitability of the earth.
The effects of acid rain, felling and burning are having drastic consequences
on the extent and structure of forests throughout the world. An important
task for the SAR carried on ERS-l is to provide information on these
processes.
Operating SAR in mountainous regions gives rise to another peculiarity
of SAR imaging, known as lay-over. Because the SAR positions objects in
the cross-track direction by the time of propagation to them, the tops of
mountains are apparently moved towards the radar compared to their
bottoms, in the two-dimensional image (see Fig. 11.16). In the extreme, as
the slope steepens, the top may be closer to the radar than the bottom. This
will cause their respective positions in the image to be reversed compared
to the normal two-dimensional map projection. The slopes on the lee-side
of mountains compared to the radar are also likely to be in radar shadow
(Fig. 11.16). Such effects hamper the use of SAR in studies of snow
mapping, but can be mitigated by the use of digital terrain models in
combination with the SAR data.
The role of SAR in ice mapping has already been mentioned. It seems
clear that SAR has a lot to offer in this application. This will include the
gathering of information on ice age and thickness, ice type and ice dynamics.
Together with location of ice leads and icebergs, this information should
help to make operations at higher latitudes safer.
SAR techniques have also been dramatically successful at imaging the
planets from orbiting spacecraft. First among these are the images of Venus
taken by the Pioneer and Venera 15 and 16 spacecraft. Venus is entirely
covered by cloud and nothing can be seen of the surface using ordinary
optical methods-indeed, the speed and direction of rotation of the planet
were not known until radar revealed it to be spinning in a retrograde fashion.
The Pioneer images showed that the surface topography is not unlike that
of earth, but with any oceans now gone and with what appears to be an
out-of-control greenhouse effect, giving high temperatures and very strong
winds. The Magellan spacecraft has now produced much higher-resolution
images of Venus, which show a remarkable amount of detail, and should
tell us more about the differences between the two planets.
SAR-type processing can also be carried out when the radar is stationary
but the target is moving. This technique, known as inverse SAR, is useful
for imaging objects, and has been used in radar astronomy to map planets
by making use of the differential doppler shifts across the planetary surface,
252 UNDERSTANDING RADAR SYSTEMS
Radar
E shadowing
Image
positioning A' 8' C' E' D' shadow G'
'----y----'
lay-over
I
Map I II
positioning A" B" C" D" E" G"
Figure 11.16 Effects that occur when SAR is used in hilly terrain. The map positioning of
points has been displayed so that point C coincides in the image and map. The distortion due
to terrain effects is obvious, as well as the more extreme effects of lay-over and shadowing.
11.8 SUMMARY
The scattering behaviour of the earth's surface at radar wavelengths can
provide useful information about many natural processes. Since most
RADAR REMOTE SENSING 253
Key equations
• The Bragg scattering condition:
Ln = n).I(2 sin 0) [m]
• The length of the synthesized antenna:
D = R).ld [m]
• Along-track resolution of a SAR:
ra = dl2 [m]
• Along-track phase history of a scatterer in the SAR beam:
./,( ) _ 2nV2 2
'l't - - - - t [radians]
)'R
11.9 REFERENCES
1. Radar Reflectivity ofLand and Sea, M. W. Long, Lexington Books, Lexington, Mass., 1975.
2. Microwave Remote Sensing: Active and Passive, F. T. Ulaby, R. K. Moore and A. K. Fung,
vols 1-3, Addison-Wesley, Reading, Mass., 1981.
3. Synthetic Aperture Radar, J. J. Kovaly, Artech House, MA, 1976.
4. Spaceborne Radar Remote Sensing: Applications and Techniques, C. Elachi, IEEE Press,
New York, 1987.
11.10 PROBLEMS
11.1 A 15 MHz HF radar is arranged to measure tidal currents across the entrance to a harbour.
The dominant Bragg line is observed to have a doppler shift of 0.7 Hz. just as a large ship is
seen approaching the harbour. Do you think it is important to convey the result of your
observation to the ship?
11.2 A scatterometer mounted on a tower, as shown in the diagram, uses frequency modulation
of bandwidth B [Hz] to give a maximum possible time resolution 1/ B [s]. Show that this gives
a slant range resolution
r, = c/(2B)
Each of these resolution cells corresponds to an independent measurement of the RCS of the
ground (see diagram). Show that the number of independent samples available for averaging is
2B
N ~ - hfJ sec (J tan (}
c
where fJ is the beamwidth and () is as shown in the diagram. Plot the variation of N as ()
increases from 10° to 70°.
RADAR REMOTE SENSING 255
11.3 Assuming a beam width of at most a few degrees, show that the annulus generated by a
pulse-limited altimeter has approximate area 2nRcr, where the satellite is at altitude R, c is the
speed of light and r is the pulse length. Show also that the annulus forms while the leading
edge of the pulse is still in the main beam if the condition
r < RfJ2 /(8c)
is met.
The ERS-l altimeter operates with a 1.2 m diameter dish at a frequency of 13.8 GHz; the
satellite has a nominal altitude of 785 km. If it were operating in pulse-limited form, show that
the pulse length would need to be less than 0.11 JlS. For a pulse length of 0.1 Jls, a loss factor
of 50 per cent and a transmitted power of 500 W, what would the signal-to-noise ratio be at
the satellite if the surface has an RCS of 5 dB m 2 m - 2?
11.4 Compare the azimuth processing requirements of (a) an air-borne SAR travelling at
200 km h - I with aim antenna, operating at a range of 50 km, and (b) a space-borne SAR
travelling at 7.5 km s - I with a 10 m antenna, operating at a range of 1000 km, if real-time
operation is required. Do the comparison for the cases where both platforms are carrying a
C-band (6cm) and an L-band (24cm) radar.
11.5 The returns from each pulse of a SAR come from a swath defined by the maximum and
minimum slant ranges (see diagram). Returns from successive pulses must not overlap on
256 UNDERSTANDING RADAR SYSTEMS
reception. Show that, for a SAR platform whose speed is V and whose antenna has an along-track
dimension d, this leads to the two conditions (using Eq. (11.16)
(a) PRF,;:; c/(2R,)
(b) R,';:; cd/(4V)
What is the widest possible swath that can be imaged by an air-borne SAR if it is carrying a
2 m antenna and travelling at 300 km h - I? What is the best possible resolution in the along-track
direction available to a space-borne SAR travelling at 7.5 km s - I if it is designed to produce
images from a slant swath 100 km wide?
11.6 The ERS-I satellite operates at incidence angles varying from 20.\" to 25.9°, with a
mid-swath incidence angle of 23". Are lay-over and shadowing likely to be major problems?
Compare this with the effects likely in SAR images gathered by a long-range air-borne SAR
operating at an altitude of 10 km and with coverage from 50 to 100 km ground range.
CHAPTER
TWELVE
GROUND-PROBING RADAR
• Applications
• Propagation in the ground
• Carrier-free radar
• Wideband antennas
12.1 INTRODUCTION
257
258 UNDERSTANDING RADAR SYSTEMS
There are many other ways of imaging the subsurface and the interior
of structures; these include seismology, ultrasound, magnetometry, impedance
imaging and low-frequency induction methods. However, ground-probing
radar systems have the advantage of being relatively lightweight, mobile,
well focused and easily configured to detect particular types of target. The
disadvantage of ground-probing radar is that good radio propagation is only
achieved in dry or low-conductivity materials, and the penetration into wet
rock or soil may be restricted to a few metres.
Over the years, a great many articles have been published on ground-
probing radar in a wide variety of journals and conference proceedings.
Fortunately, in 1988, a special issue of the lEE Proceedings-F was devoted
to this topic! , and the introductory article presents a comprehensive review
that is a useful starting place for those wishing to discover more.
10
Ice
Sea ice
E
-
I g
e 2 X 102 iii
CQ
~
~
c e
.2 8
~
C
::l ....
2 '"Co
~ '"
'"
;>, Ice ..Q
'"~ 10- 1
Sea ice 20 ;...
'"
~
V
c Permafrost
6
0 ~
f-
lO-31~M--H-z-------l~M~H~Z--------1~OO~M~H~z------~1-G~H~z--~
Frequency
100 MHz and buried object location to below 1 GHz. Few systems operate
above 1 GHz and antenna constraints effectively prevent operation below
1 MHz.
The range/depth resolution is determined by the bandwidth, but this is
not always the bandwidth of the transmitter (which in the case of impulsive
radar may be inherently very wide ). One limiting factor is the transfer function
of the antenna and, although wideband antenna designs are used, the
transmitter pulseshape has usually suffered some distortion before it enters
the ground. A second limitation is the nature of the ground, which acts like
260 UNDERSTANDING RADAR SYSTEMS
Volts
~Tim'
100
Exponential decay
25 ns
26
Time [ns]
speaking, below 1 GHz, this does not present too many engineering diffi-
culties and commercial units are available. Sampling the receiver output is
much more of a problem. With spectral components up to 1 GHz, direct
AID conversion would be prohibitively expensive with today's technology,
and some method of slowing down the data rate is needed. One method
commonly used amounts to a 'stroboscopic' or aliasing technique; the echo
pulse from a given range gate is successively sampled a few nanoseconds
later on each PRF sweep so that the shape of the echo pulse is gradually
revealed, as shown in Fig. 12.4. Some time-domain refiectometers work in
this way. There is no inherent integration gain in the stroboscopic process,
and the SNR is less than for direct sampling. However, since subsurface
targets are stationary, there is plenty of time to integrate after stroboscopic
sampling, provided the radar is not on a fast-moving platform.
An example of a recording made by a relatively simple carrier-free radar
with a stroboscopic receiver is shown in Fig. 12.5a. Physical investigation
of the ground confirms that the major features have been revealed correctly,
as shown in Fig. 12.5b.
There has been some military interest in the use of carrier-free radar
technology. Irregular wideband impulses are not easily detected and located
by ESM (electronic support measures) receivers (see Chapter 14), and so
these radars have some immunity to countermeasures. It is also difficult to
Transmit Echo
pulse
.. iILn
100 ns
(a) x[m]
(b) xlm]
100 200 300 400
2
table
E 4
.c Sand
.,
15..
""'"+-- Permafrost Lithological
o 6 boundary
Figure 12.5 (a) Carrier-free radar profile of the ground made from a helicopter at a height of
50 m and travelling at 50 km h - I (after Finkelstein at the Riga Institute of Civil Aviation,
Latvia). (b) Geological investigation confirms the observations.
Angle principle If the dimensions of a simple antenna are doubled and the
wavelength is also doubled, the performance remains the same. This means
that the impedance, polarization, antenna pattern, etc., are invariant to a
change of scale in proportion to wavelength 4. If the shape of the antenna is
determined entirely by angles, the performance has to be independent of
266 UNDERSTANDING RADAR SYSTEMS
Time III
o x
o x
o x
o x
frequency. In practice, there is an upper and lower limit, but between them
a very large bandwidth can be achieved. The most commonly used designs
are planar and helical spirals, and log-periodic structures.
Horns At the higher frequencies, horn antennas can be used, and the
bandwidth may be extended by careful design of the shape, including internal
ridges, and by dielectric loading.
With some of these antenna designs there is an effect in which the active
region of the antenna moves to different parts of the antenna as the frequency
is changed, causing a frequency-dependent delay. Further details can be
found in many antenna books, and for an interesting history of the
development of frequency-independent antennas try Chapter 7 of Weeks 5 •
(a)
~r x
(b)
Air
.1".';
Figure 12.7 Response of a buried scatterer: (a) geometry showing apparent increase of depth
with horizontal distance; (b) characteristic hyperbolic shape observed. In practice, the picture
is more complicated, especially for elevated antennas, because of diffraction at the surface and
interaction between space-wave and surface-wave effects.
different ranges by a moving radar, because of the finite width of the antenna
pattern (see Fig. 12.7a). The actual range is given by
2
R = -J(x 2 + d 2 ) Em] (12.3)
V
12.6 SUMMARY
12.7 REFERENCES
I. Special issue on 'subsurface radar', lEE Proc.-F, 135(4), 1988. [A very useful collection
of papers.]
2. Nonsillusoidai Ware/or Radar and Radio Communication, H. F. Harmuth, Academic Press,
New York, 1981.
270 UNDERSTANDING RADAR SYSTEMS
12.8 PROBLEMS
12.1 An article in the special issue of the lEE Proceedings":'F on subsurface radar describes an
impulse radar for mapping underground utility lines 6 . The 3 ns transmitter pulse has an
amplitude of 200 V and the receiving system uses a preamplifier and a sampling oscilloscope.
If the noise level at the sampling head is 10 m V, calculate the minimum gain of the preamplifier
if objects are to be detected to a depth of 4.3 m. Assume a round-trip attenuation of 23 dB m - I
and 17 dB losses, mainly in the antenna system. Remember that voltage ratios are converted
to decibels by 20 log ( VI / Vz ).
12.2 What is the maximum unambiguous PRF that could be used if the radar in problem 12.1
were to be instrumented for a maximum range of 4.3 m? Assume the phase velocity of radio
waves in the ground to be c/3.
12.3 If the system in problem 12.1 were to pass over a metal grate in the road, what do you
think would happen?
12.4 If the radar in problem 12.1 were to be used for probing very dry rock, such as on Mars,
how much greater might the penetration into the ground be? Use Fig. 12.1 to estimate the
attenuation.
12.5 If a low-frequency version of the radar in problem 12.1 were designed specifically for
geological applications and operated at a frequency near 10 MHz, what depths might be probed?
12.6 Calculate the maximum unambiguous PRF and the likely antenna size for the radar in
problem 12.5.
CHAPTER
THIRTEEN
MULTISTATIC RADAR
13.1 INTRODUCTION
In a monostatic radar system the transmitter and receiver are located at the
same place, sometimes sharing a single antenna. If the transmitter and receiver
are widely separated, by a baseline typically one-third of the distance to the
target, then the system is said to be bistatic. If there are several widely
distributed receivers associated with a single transmitter, or (more rarely)
several transmitters, then the system is multistatic. Bistatic radar is thus a
subset of multi static radar, but is by far the most common form used.
The earliest radars had to be bistatic since it was not practicable to build
pulse waveform radars and T -R switching. An early example of a bistatic
radar was built by Dr Albert Taylor and his assistant, Leo Young, at the
US Naval Research Laboratory in 1922. Using a 500 Hz CW modulated
transmitter, and a receiver placed in a car, Taylor and Young drove across
the Potomac River and detected a variety of targets, including a wooden
ship (for more details see Swords!).
271
272 UNDERSTANDING RADAR SYSTEMS
Figure 13.1 shows the basic arrangement of a muItistatic radar system. The
transmitter illuminates a sector either by a scanning pencil beam or by
floodlighting the whole area. Usually the transmitter plays no further part
in the radar, although it could be part of a monostatic radar in its own right
(as was Chain Home during World War II).
The first signal detected by the receiver is the direct pulse, which has
travelled along the baseline from the transmitter. This signal initiates the
receiver timing, from which the target range will be determined. The receiving
antenna system uses either a floodlight beam, a staring array of many fixed
beams, or a scanning pencil beam carefully synchronized to a scanning
transmitter beam.
Floodlight illumination - - - f - -
-'\-_--+-- Scanning beam
Scanning beam - - - ' - f - - - :
·,......\-~)tarml!. array
Direct
Transmitter pulse Receiver
Figure 13.1 Basic antenna arrangements used by bistatic and multistatic radar systems.
MULTISTATIC RADAR 273
radar sites (two aircraft, two pieces of coastline, etc.) genedllly cost more
than a slightly larger single site.
• A further complication in using two sites is that both must be able to
see the target. For low-flying targets this increases the problem of terrain
obscuration.
• There is usually a 'dead zone' between the sites in which it is difficult
to detect targets; see Sec. 13.5.
(a)
r - - I - - - - - - Surveillance area
Two-way sidelobe
protection
Tx.Rx
(b)
->r~'----'c-----Area under
surveillance
.x:=.d,----Arleas of one-way
side lobe protection
oo......---"''''':--''r--\-----Area of
two-way side lobe
protection
Tx Rx
Figure 13.2 (a) The clutter in the side10bes of a monostatic radar antenna is attenuated on
both transmit and receive. (b) With a bistatic system, there are clutter patches illuminated by
the sidelobe of one antenna and the mainlobe of the other.
MULTISTATIC RADAR 275
Using the same notation as in Chapter 1, the power flux density arriving at
the target is
. PtGtL
Flux densIty = ----t ( 13.1 )
4nR~
SOLUTION Assuming all other system factors are equal, the R4 losses
are the important factor. For the monostatic case
R4 = 10 20 [m4] == 200 [dB m4]
and for the bistatic case
R~R; = 10 18 [m4] == 180 [dB m4]
In the second case, a transmitter of only 1/1 OOth the power of the
monostatic radar is required for the same SNR to be received. The
bistatic receiver is relatively safe from attack because it is electronically
silent, and the transmitter is safer because the transmissions are weaker
and consequently harder to detect.
(a)
(b)
e>l
e»l
Figure 13.3 (a) Surface of constant SNR for a bistatic system. (b) The intersection of the
surface with a plane gives the ovals of Cassini.
MULTISTATIC RADAR 277
The first piece of information used for target location is the time delay between
the transmission of a pulse and the reception of an echo (analogous to range
in a monostatic radar). After finding out how to use this, we go on to examine
the different possible geometries that can be used for target location.
The range measured by a monostatic radar defines a circle (an iso-range
contour) centred on the radar, upon which the target must lie (see Fig. 13.4a).
In the case of a bistatic system, the total path length (R\ + Rr) defines a
three-dimensional elliptical shape shown in Fig. 13.4b. On any given flat
plane you can construct the appropriate ellipse by sticking two pins into a
sheet of cardboard, tying a piece of cotton between them of a suitable length
to represent (R\ + Rr ), and drawing out the shape with a pencil while keeping
the cotton tight, as in Fig. 13.4c.
The importance of the elliptical path shown in Fig. 13.4b is that usually
Rl and Rr cannot be measured independently because the transmitter~receiver
timing information only gives the combined distance 2r = (R 1 + Rr)' Thus
we know only the total length of the cotton, and not the two separate
distances from the pins to the pencil. Note that these iso-range ellipses are
not coincident with the constant SNR ovals of Cassini. A target flying around
a constant-range ellipse would thus have a different SNR at each place on
the contour; this is not true monostatically.
If only time-delay information were available, we could draw the
appropriate ellipse, but we would not know where the target lay on that
ellipse (in the same way, knowing only the range in a monostatic radar puts
the target somewhere on a circle). Angular information is needed to make
further progress, and there are several ways in which this may be introduced.
278 UNDERSTANDING RADAR SYSTEMS
(a) (b)
(c)
Elliptical Geometry
This is the obvious, and most common, method of target location (see
Fig. 13.5a). It is necessary to know: 2a, the baseline; 2r = (R( + Rr ) the
round-trip distance; and Or' the target azimuth, measured at the receiver.
The range Rr of the target from the receiver can be found by using the cosine
formula to solve the triangle in Fig. 13.5a, giving
r2 _ a2
R =---- Em] (13.6 )
r r - a cos Or
Knowing the range and the azimuth from the receiver, the target position
is found in the same way as for a monostatic radar. The target height can
be determined by introducing the elevation angle in a similar way.
MULTISTATIC RADAR 279
(a) Target
Tx 2a Rx
(b)
(c)
Tx Tx
Figure U.S Methods of target location: (a) elliptical geometry; (b) theta-theta geometry;
(c) hyperbolic geometry.
when the transmitter is also a monostatic radar in its own right and is
measuring the transmitter azimuth. This method can also be employed if the
target jams the radar; in this case the radar transmitter is turned off and
the target is tracked passively from the two angular measurements made on
the jamming signal.
Theta-theta target location requires: 2a, the baseline; and et , en the
target azimuth measured at the transmitter and receiver.
Note that, as Fig. 13.5b shows, there are places near the baseline where
the accuracy is poor. It is generally true to say that the advantages of
multistatic radar are maximized when the target is at intermediate range and
near-broadside to the baseline.
Hyperbolic Geometry
This is used with two transmitters and one receiver. Here the target is located
from the time d(fference between the two transmitter signals arriving at the
receiver, which places the transmitter on a hyperbola. The receiver beam is
then used to fix the position on the hyperbola, as shown in Fig. 13.Sc. The
information needed is: 2a, the baseline; 2r, the difference in the round-trip
distances; and en the azimuth of the target, measured at the receiver.
From these measurements, the target can be located from the azimuth
and the distance to the receiver given by
a 2 _ r2
R =---- Em] (13.7)
r r + a cos er
A comprehensive table of target location equations, extended to three
dimensions, is given by Skolnik 2.
(a) (b)
Tx.Rx
Figure 13.6 Zero isodop contours for (a) monostatic case and (b) bistatic case.
Target
Tx Rx
Figure 13.7 Geometry for evaluating the doppler shift.
Since the range does not change, there is no doppler shift and the aircraft
would be hidden by ground clutter (although it would not pose much of a
threat). In the case of a bistatic system, the iso-range ellipses represent the
general zero isodops, but there is also the particular case of an aircraft flying
along the baseline from the transmitter to the receiver; as R t increases, Rr
decreases by the same amount, and (as there is no net change in the total
path) we see zero doppler again. In general, the doppler frequency represents
a component of the target motion along a line roughly bisecting the bistatic
angle 21/1 as shown in Fig. 13.7.
When the transmitter and receiver platforms are moving, as in the case
of air-borne or ship-borne radar, the doppler frequency is modified by the
component of platform velocity along R t and R r • The best way to express
this mathematically is to define unit vectors U t and U r along R t and Rr
respectively, such that the doppler shift becomes
[Hz] (13.9)
282 UNDERSTANDING RADAR SYSTEMS
One must not become too gloomy about the complications posed by
bistatic doppler and the problems of doppler spreading of clutter. First, this
is a subject amenable to computer simulation and prediction. Secondly, the
transmitter and receiver platform speeds can be arranged for the benefit of
the radar user. An example of this is clutter tuning, which concerns air-borne
detection of surface vehicles such as tanks; normally this is a difficult task
for monostatic radar because the high platform speed of the aircraft and the
finite width of the antenna beam causes a significant clutter spreading
problem (see problem 3.5 at the end of Chapter 3). Using a bistatic
arrangement, the speeds of the transmitter and receiver aircraft can be
MULTISTATlC RADAR 283
adjusted to give zero platform doppler shift over the area of interest and the
capability of being very sensitive to slowly moving land targets. This
procedure is known as engaging complementary motion.
13.6 APPLICATIONS
Pulse --~-----'~~'-,,~
--·Staring· array
of beams
Tx Rx
receiver so that it can keep watch on the target while following an optimal
trajectory that carries it outside the radar beam. Further details are given
in Chapter 3.
All these applications require bistatic and multistatic tracking algorithms
to be developed. Transferring the measurements onto cartesian coordinates
means that non-linear filtering is necessary, which in turn means more
computer processing power and complexity. Ifthe target is detected by more
than one receiver simultaneously, then tracking accuracy can be improved
by integrating all the data into the filter.
13.7 SUMMARY
Bistatic and multistatic radars differ in a number of ways from equivalent
monostatic systems. Some of the differences are unwanted and merely add to
the system cost and engineering complexity, but others can be exploited to
produce significant operational advantages.
Perhaps the biggest single advantage of multistatic operation is the
reduced vulnerability of the receiver to jamming or physical attack by
anti-radiation missiles. Another big advantage, in the case of mobile systems,
is that clutter tuning may be employed to increase the sensitivity of the
system to slowly moving targets. A higher PRF can be used with bistatic
radar than with monostatic, which also aids sub-clutter visibility.
The original reason for the application ofbistatic techniques in the 1920s,
which was to isolate the receiver from the transmitter when CW waveforms
were used, remains valid today and most large HF over-the-horizon radars
have transmitters and receivers separated by over 100 km (see Chapter 10).
These radars also follow the typical bistatic format of a single transmitter
beam and several narrow receiver beams. The cost of multi beam receiving
antennas, and other system complexities, means that multistatic radar will
never replace monostatic radar in general usage, but in certain applications
it remains a powerful technique.
Key equations
• The bistatic radar equation:
p = PtGtGrO"b}·2LtLr [W]
r (4n)3R?R;
13.8 REFERENCES
I. Technical History of/he Beyinnin!Js 0/ RADAR, S. S. Swords, Peter Peregrinus for the lEE,
Stevenage, Herts, 1986.
2. Radar Handhook, Ed. M. E. Skolnik, McGraw-Hili, New York, 1990.
13.10 PROBLEMS
13.1 A bistatic short-range air surveillance radar operates at a frequency of 3 GHz. An aircraft
flies directly towards the transmitter at 100m s - 1 •
(a) At the peint where its path brings it at right angles to the receiving beam, what
doppler shift is recorded?
(b) If the transmitter is a monostatic radar in its own right, what doppler shift would it
detect?
(c) Explain the difference.
13.2 Assume that the radar in problem 13.1 has a baseline of 30 km. An echo from a target
arrives at the receiver at an angle of 30" to the baseline and 200 lIS after the direct pulse from
the transmitter arrives.
(a) What is the round-trip distance 2r?
(b) Calculate the range of the aircraft from the receiver.
(c) Calculate the range of the aircraft from the transmitter.
Neglect the curvature of the Earth and the aircraft height and treat this as a two-
dimensional problem.
13.3 (a) In the problem above, if the target were travelling parallel to the baseline at 100 m s - "
what doppler shift would be recorded?
(b) Calculate the bistatic angle.
13.4 A I GHz, 1 W transmitting buoy is dropped in a shipping channel s!lch that ships must
pass within I km of it. Overhead, an aircraft is listening for the echoes with an antenna of gain
20 dB. Assume that the buoy has no useful antenna gain (assume a gain of unity) and the
aircraft receiver requires a signal strength of - 130 dB W for detection.
(a) What is the maximum height at which the aircraft could detect a ship of bistatic RCS
30 dB m 2 (neglect losses).
(b) Woald the air-borne receiver have a dynamic range problem with the direct signal
arriving only a few microseconds before the echo?
(c) If the aircraft had its own monostatic radar on board, would this need more, or less,
power than the transmitter on the buoy?
CHAPTER
FOURTEEN
ELECTRONIC WARFARE
The jamming of radar systems has been going on since radar was invented,
and is now very sophisticated.
A modern war is won by the side that best exploits the electromagnetic (EM)
spectrum. During the Gulf War in 1991, Iraqi radars were blinded by
electronic countermeasures and the opposing Allied forces were able to gain
control of the air. Once air superiority has been achieved, victory is often
just a matter of time. One of the key elements in a modern battle is how
well the attacking and defending radars of each side stand up to the
electromagnetic onslaught from the other.
The objectives of electronic warfare (EW) systems are simple:
286
ELECTRONIC WARFARE 287
~
Escort jammer Stand-off
jammer
Self-screening jammers
Expendable
jammer
Putting in values:
Pt = 53 dB W
GtGr = 72 dB
a = 10 dB m 2
),2 = -12.7 dB m 2 ), = 0.231 m
Lsys = -7.0 dB Typical
1/(4n)3 = -33.0dB
I/N= +145.8dBW- 1 3 fJ.s pulse, 3 dB noise figure
I/(SNR) = -13.0dB
R!ax = 215.1 dB m 4
Rmax = 238.5 km
and so
Pr = -126.7 dB W
(411:)2 = 22.0 dB
R2 = 107.5 dB m 2 Aircraft range of 238.5 km
I/GI = 0 dB
I/Gr = -36 dB
1/..1? = 12.7 dB m- 2
llLlotal = 7.0dB
performance of radars. Provided one knows exactly how the enemy radar
works, it is relatively easy to defeat it. A good guide on the use of ELINT
to analyse enemy radar is given by Wiley 3. The key to ECCM is for a radar
not to be detected in the first place and to have a low probability of intercept
(LPl). There are many ways of achieving LPI, but they are all based on
spectral spreading of the radar signal by such techniques as FMCW, long
pseudo-random codes, etc. Radar systems can accurately compress these
signals on reception to a narrow final bandwidth. The intercepting ECM
receiver, not knowing the compression algorithm, must receive these signals
over the full swept bandwidth and so suffers a large noise penalty. Another
defence against ELINT and ECM is to use separate peacetime and wartime
operational modes, and to transmit deliberately misleading surveillance-type
signals during peacetime to try to get the enemy to develop the wrong type
of jammer.
One final deception system that should be mentioned here is the use of
decoy RPV aircraft. These are relatively inexpensive and can be flown in large
numbers at the enemy until their air defence weapons are exhausted. The
possibility that the enemy radar can recognize the decoys and ignore them
can be countered by placing a nuclear or biological weapon on the occasional
one to ensure that they must all be treated seriously.
The ECM-ECCM game can be played indefinitely. There are many
sophisticated modifications that can be made to radar transmitter waveforms,
antenna patterns and receiver techniques to defeat denial and deception
jamming. However, it is probably also true to say that the countermeasures
to a radar are always less expensive to build than the radar itself. The final
arbiters in electronic surveillance are probably, therefore, espionage (to know
how the countermeasure will work in wartime) and money (to fund the
development of a counter to it). All of this expensive game can be invalidated
if targets are not visible to radar in the first place; this has been the driving
force behind the Stealth programme.
'Flying Wing' and the Avro Vulcan bomber showed that the aircraft shape,
as well as the construction material, was important in the reduction of the
RCS.
A change in tactics also occurred after the war; the large defensive
bomber formations, and escorting fighters, were replaced by solo nuclear
missions in which single bombers were required to navigate through enemy
defences alone. Escaping enemy ground fire by flying at very high altitudes
was eventually abandoned after the U2 of Francis Gary Powers was shot
down in 1960. Flying at low altitudes was dangerous because of the accuracy
of radar-controlled anti-aircraft artillery, unless the aircraft flew very low to
get beneath the radar defences. The solution lay in adopting an idea tried
in World War I, to make aircraft 'invisible' by covering the airframes in
translucent materials, and then extending the concept to include the radar
spectrum (see Sweetman 4 , for example).
The maximum radar detection range (Eq. ( 1.15 )) can be summarized as
[m] (14.1 )
The RCS u must be changed dramatically if the radar performance is to be
seriously affected. Computer modelling of airframes and ship structures
shows that most of the scattering comes from flat surfaces and corner
reflectors, formed where different parts of the structure join together.
Table 2.1 gave the RCS of flat plates and corner reflectors; try putting a few
numbers in to get a feel for the values that can occur. The RCS of most
structures can be dramatically reduced by choosing an inherently low RCS
design and then using smooth, continuously curved shapes and avoiding any
right angles. In the case of an aircraft, this avoidance of right angles may
mean that a single fin cannot be used and the aircraft must be designed
without one (the flying wing approach used in the B2 bomber) or with two
inclined fins (as used on the F 117 fighter).
After shaping, any remaining 'hot spots', such as the reflections from
engine air intakes on aircraft, can be covered with radar-absorbing material
(RAM). RAM is generally some form of lossy dielectric material built into
a wafer, which is tuned for maximum absorption in a particular radar band.
Care also needs to be taken to ensure that the radar antennas on ships and
aircraft do not act as RCS hot spots.
In a battIe scenario, Stealth is a very effective counter to radar, especially
when other ECM activities are going on to reduce the radar efficiency. There
are, however, a few antidotes to Stealth. Wideband carrier-free impulse radar
may be used to try to exploit RCS weaknesses in the radio spectrum (see
Chapter 12) and, in general, radar needs to move to lower frequencies to
combat Stealth. Over-the-horizon radar (Chapter 10) may be effective
because the scattering mechanism is more related to radar-induced currents
flowing throughout the structure of the target than from the individual
scattering facets on the surface that dominate the RCS at microwave
294 UNDERSTANDING RADAR SYSTEMS
14.5 SUMMARY
The very success of radar at detecting and tracking military targets means
that it has become the target of electronic attack itself. As soon as a military
radar system is turned on, its EM emissions will be detected unless it uses
spread spectrum LPI techniques to evade detection. Any radar signals
detected will be analysed by ESM receivers and an ECM strategy formulated;
this may include the use of Stealth to reduce the RCS of the target. While
electronic countermeasures are effective, and expensive to defend against
using ECCM techniques, the balance at present probably resides on the side
of radar, provided the waveforms are controlled intelligently.
14.6 REFERENCES
14.7 PROBLEMS
14.1 Consider a radar speed trap based on a 10 GHz 'unity' radar system (I W, unity antenna
gain, zero losses, I JlS pulse). What is the SNR for a single-pulse detection of a car at a range
of 100 m if the car has a radar cross-section of 10 m 2 ?
14.2 If the motor car in problem 14.1 contained a spot jammer, how much would this need to
increase the noise level in the radar receiver to reduce the range to 10 m for the same received
SNR?
14.3 What noise power would the jammer have to radiate to achieve the result in problem 14.2,
assuming the noise was all concentrated in the same bandwidth as the radar signal?
14.4 If the radar trap above were to use random frequency hopping over the range 10- 30 GHz,
and the jammer were to counter by barrage jamming the entire band, by how much would the
total jammer output noise power have to be raised?
ELECTRONIC WARFARE 295
14.5 If. as an alternative. the radar in problem 14.1 were to attempt to burn through the spot
jamming using (perfect) coherent integration. how many pulses would need to be integrated?
14.6 Suggest a suitable PRF for the radar and estimate whether the integration time determined
by problem 14.5 would give adequate doppler resolution to book a motorist for speeding.
14.7 What would seem to be the most economical strategy to avoid a fine for speeding:
(a) to build an in-car radar warning receiver;
(b) to develop a radar jammer;
(c) to cover the car in Stealth materials?
CHAPTER
FIFTEEN
RECENT DEVELOPMENTS
15.1 INTRODUCTION
In this chapter we look at current radar technology and find out how it is
being used. We will try to predict the future of radar in the next chapter.
The antenna has been the subject of some of the most interesting radar
developments in recent years. Domestic satellite dishes are representative of
the old level of antenna technology because they must be mechanically
pointed at their target. Radar versions of these parabolic reflectors are rotated
by heavy and expensive (but reliable) turning gear, and a target can only
be observed once every few seconds when the dish faces that direction.
At the centre of a reflecting antenna, such as a satellite dish, is the device
that illuminates it, called the feed. Feeds can be dipoles, microwave horns
or other devices for radiating or receiving electromagnetic waves. The purpose
of the reflector is merely to increase the collecting area, and hence the gain,
of the antenna. However, a large collecting area can also be achieved in an
entirely different way. It is possible to arrange many dipoles on a flat plate
296
RECENT DEVELOPMENTS 297
Broadside
lr
N
Figure 15.1 A linear array of dipole elements.
In optics, the field arising from a diffraction grating has two components;
the diffraction pattern from a single slit and the interference pattern created
by the interaction of coherent signals from different slits. It is exactly the
same with antenna arrays. Each element of the array acts as a source with
its own diffraction pattern E 1 (0), and the radiation from the sources interacts
to form an interference pattern, which is usually called the array factor (AF)
given by
- Ea((J) -
AF-~-- J N (sin[Nn(djJ.)SinO J) [ J (15.3 )
Ed8) N sin[n(dj}.) sin OJ
The radiating elements are usually simple devices such as dipoles, slots,
patches or sometimes small horns. All these elements have fairly wide
diffraction patterns and the antenna pattern is determined by the much
narrower array factor. Even so, care must be taken not to steer the array
beam outside the pattern of the individual radiating elements. A normalized
array pattern is shown in Fig. 15.2 in comparison with the diffraction pattern
of a single half-wave dipole.
Part of the purpose in rearranging Eq. (15.2) to get Eq. (15.3) is because
the term in large parentheses in Eq. (15.3) can be shown to have a maximum
value of unity. This means that the field strength has been increased by N J
times the field from a single element, and therefore the power gain of the
array has been increased by N. Some other properties of array patterns are
now outlined.
The half-power beanmidth of the pattern steered off-broadside through
an angle 8 can be found from Eq. (15.3) by setting the expression in large
RECENT DEVELOPMENTS 299
/
/
/
'-'-'-/
,
/
/
,,
/
/
,,
/
/
/
, \
/ \
/ \
/ \
/ 0.4 \
/ \
/ \
/ 0.3
\
/ \
/ \
/
e [degrees]
I I [ I I [ [ I [ I I •
o 0.2 0.4 0.6 0.8 0.9 0.99 1.0
sin e
Figure 15.2 The normalized array factor for a seven-element array, of spacing d = i./2,
compared with the diffraction pattern of a single element.
beam) is about 51 for the uniform illumination assumed in Eq. (15.6), but
is between 60 and 70 for tapered distributions, according to the sidelobe
level required. The beamwidth factor is also higher for apertures that are
round, rather than rectangular.
Nulls occur in the pattern at angles 0nuJl given by
sin 0nuJl = ±kA/Nd k = 1,2, ... [ ] (15.7)
Often, element spacings near )./2 are used, giving only one mainlobe (plus
one to the rear, if the array is' not mounted above a reflecting plane). For
wider element spacings, grating lobes (other principal maxima, or 'secondary
beams') occur when
sin Ogl = ±H/d k = 1,2, ... [ J (15.8 )
With careful design, the individual element pattern can sometimes be used
to null out grating lobes near 90°, as shown in Fig. 15.2.
Side lobes (secondary maxima) occur at
. 0 (2k + 1»),
sm 51 = + -----'- k = 1,2, ... [ J (15.9)
2Nd
The first sidelobe has a power gain of - 13.4 dB with respect to the peak
gain of the array. This is a well known result arising from a rectangular
illumination function. Those of you familiar with Fourier transform (FT)
theory will realize that the FT of a rectangular function (in this case the
aperture) gives rise to a (sin x )/x function of a similar form to the array
factor. Equation (15.3) can be approximated, for small values of 0, by an
expression of the form
( 15.10)
Thus the far-field pattern of the array can be represented by the FT of the
spatial illumination function, provided that the element spacing is close
enough to avoid grating lobes. For this reason, FTs are used extensively in
antenna theory.
Some of you may be confused at this point because you have only used
the FT to transform between the time domain and the frequency domain.
Table 15.1 shows you how to relate this new way of using the FT to the
more usual time/frequency form. Bracewell! is particularly good at covering
this use of FTs. These days, there are some very readable books on
antenna theory, and on arrays in general. If you wish to explore further, try
references 2-4, especially 4.
Table 15.1 Use of Fourier transforms in antenna theory
Waveforms Antennas
~
*
X(w): T sin (wT/2) E(sin Il) '" E1(sin Il)sin INTr(d/)..) sin III
(wT/2) NTr(d/)..) sin H
Target
v
Oil-rig
Figure 15.3 An air-borne surveillance radar and antenna pattern showing sidelobe structure.
RECENT DEVELOPMENTS 303
Normal
Beam
r
Figure 15.4 Beam steering by the introduction of a progressive time delay between array
elements.
e [degrees]
e[degrees]
l
Grating lobe
l
Figure IS.s The beamshape for a five-element array with d = A.: (a) the broadside pattern;
(b) steered 50° off-boresight.
306 UNDERSTANDING RADAR SYSTEMS
For the problem above, this gives a spacing of near ), for a broadside beam,
near ), / 2 if the beam is steered to ± 30° and near ), / 4 if the beam is to be
steered to ± 60°. When weighting is used, a slightly smaller element spacing
is needed and a suitable approximation is given by Radford 6 :
d
<
.
I.
(I
1+ sin 8s -
1.5)
N Em] (15.14 )
Diodes
Ferrites
Signai--+-.......-l
Control
Waveguide
Beam ports
I
Low-noise pre-amp I
I
I
Image rejection mixer ·~FirstLO
I
I
IF filter I
I
IF amp I
I
-'
T
Mixer Second
LO
Video amps
Sample and
hold
T f
where v" = complex signal from nth channel, w,. = weighting coefficient,
e - j21tn(dl).) sin 8 = steering phaseshift and Cn = correction factor. Corrections
are necessary for several reasons. These include errors in the position of the
element, temperature effects and the difference in behaviour between those
elements embedded in the array and those near the edge, caused by coupling
effects with neighbouring elements.
Digital beamforming is now used extensively for receiving arrays, where
sufficient funding is available, and it can be found in use from small HF
sea-sensing systems up to large-scale surveillance radars. An example of a
system in service is the Patriot radar system, described in Chapter 3.
PIN diode
Power amp receiver
- few watts protection
Receiver
Driver pre-amp
Phaseshifter Phaseshifter
r, r------:l
I/\-./..----'''DIA ~ A~dress for I
L -,~L!~~ ele~~~J
To other
elements
Signal
in/out
to beamformer
Given the flexibility of phased array radar, the traditionally separate roles of
surveillance, tracking and even radio communications can be taken over by
a single multifunction radar system. The system's resources (beams, receivers,
computers) must be allocated efficiently to carry out all the required
312 UNDERSTANDING RADAR SYSTEMS
When several monostatic radars can see the same target, as shown in
Fig. 15.9, their outputs can be combined to improve the tracking. This method
Flight
........... path
Figure lS.9 When the coverage of surveillance radars overlaps, their outputs may be combined
to improve the overall tracking performance.
RECENT DEVELOPMENTS 313
Modern radar has gone beyond 'radio detection and ranging' to incorporate
target classification and recognition algorithms. These facilities require
improvements in resolution to identify distinguishing features on targets. We
have seen throughout this book how the resolution depends on the system
parameters, but we have not yet mentioned the contributions that can be
made by developing the hardware, the signal processing and the data
processing software.
There are many methods of increasing the resolution of radar systems.
Often, techniques were developed in other fields of research and adapted for
use with radar. Some of the current approaches involving changes in hard-
ware, or data collection strategies, are discussed below.
Improved doppler spectrum analysis There are many modern spectral esti-
mators, such as the maximum likelihood method, maximum entropy method,
minimum eigenvector method, etc. These can be used to improve the doppler
resolution and to identify targets by resolving such features as turbine blade
rotation, or the pitch, roll and yaw of an aircraft or ship. Improved spectral
estimators are also applied to phased arrays to improve the angular
information.
Super-resolution Even when the shape of a target is smaller than the range
resolution, some information may be recovered from the echo by carrying
out a deconvolution process known as super-resolution, provided that the
'point spread function' (= impulse response) of the system is known.
Essentially, the radar bandwidth is artificially increased by restricting the
possible choice of target shapes.
15.8 SUMMARY
Key equations
• Positions of sidelobes:
(2k + 1»),
sin 8s1 =± 2Nd []
15.9 REFERENCES
I. The Fourier Transform and its Applications, R. N. Bracewell, McGraw-Hili, New York, 1986.
2. Antennas, F. R. Connor, Edward Arnold, London, 1972 [Short and very readable.]
3. The Handbook of Antenna Design, vol. 2, Eds A. W. Rudge, K. Milne, A. D. Olver and
P. Knight, lEE Electromagnetic Wave series, Peter Peregrinus, Hitchin, Herts, 1983. [Long
but still readable.]
4. Radio Wave Propagation and Antennas, J. Griffiths, Prentice-Hall, Englewood Cliffs, NJ, 1987.
5. Jane's Radar and Electronic Warfare Systems 1989-90, Jane's Information Group, London,
1989.
6. Electronically scanned antenna systems, M. F. Radford, Pro£". lEE, 12S( II R), 1100-1112,
1978. [A particularly lucid review.]
7. Target Adaptive Matched Illumination Radar, D. T. Gjessing, lEE Electromagnetic Waves
series 22, Peter Peregrinus, Hitchin, Herts, 1986.
8. The MU radar with an active phased array system: I. Antenna and power amplifiers,
S. Fukao, T. Sato, T. Tsuda, S. Kato, K .. Wakasugi and T. Makihira, Radio Sci., 20(6),
1155-1168, 1985.
15.11 PROBLEMS
15.1 The Japanese MU (middle and upper atmospheres) MST radar uses an active antenna
array composed of 475 crossed Yagi antennas (each having a gain of 7.24 dB) arranged on a
circular plane of diameter 103 m. Each antenna has its own solid-state transmit-receive module,
318 UNDERSTANDING RADAR SYSTEMS
with a peak output power of 2.4 kW (see Fukao et al. S ). The operating frequency is 46.5 MHz.
What is the antenna gain and beam width ?
15.2 The beam in thl? above can be steered electronically up to 30° off-vertical. If the elements
were laid out on a rectangular grid, what would be a suitable array spacing? (The authors 9 tell
us that 'neither spatial nor electrical tapering is incorporated '.)
15.3 The triangular lattice arrangement of antenna elements (see solution to 15.2) has a
maximum separation of 1.2), in some directions. Where might the first grating lobes be expected
to appear?
15.4 We are told that it is possible to steer the beam in each inter-pulse period, which may be
as short as 400 itS. Would you expect the necessary phaseshifts to be calculated each time?
CHAPTER
SIXTEEN
THE FUTURE OF RADAR
16.1 INTRODUCTION
Has radar been developed as far as it is going to go, or does it have a bright
future with many further applications and capabilities to be discovered? As
with any branch of science or engineering, the way to answer this type of
question is to consider whether the system is limited by physical laws, by
cost or by technology.
The laws of physics applying to radar are well understood, but do not
yet limit our systems. We cannot escape the range resolution being limited
by the effective bandwidth of the system, for example, but we can apply
processing power to such techmques as super-resolution to extend the
effective bandwidth.
Cost is a serious limitation to what can be achieved; there are many
cases where phased array radars are desirable but mechanical turning gear
is used instead because it is less expensive. However, the influence of large
military research programmes has continued radar development, and the
319
320 UNDERSTANDING RADAR SYSTEMS
N=kToF -
t (1) FkToO
=---
T 6.0 6.</>
[W] ( 16.3 )
Putting Eq. (16.3) into the radar equation, and replacing Gt by 4n I (6.0 6.</> )
and Gr by 4nAe/..1. 2 gives
[ ] (16.4)
THE FUTURE OF RADAR 321
[ ] (16.6 )
(PRF) x r
and the maximum number of doppler filter channels n3 is set by the number
of pulses in the dwell time as
n3 = (PRF) x t [] (16.7)
Combining Eqs (16.5) and (16.7) with Eq. (16.2) gives the total number of
resolution cells n to be searched as
Q 1
= -- x x (PRF) x t
dO dl/> (PRF) x r
= Tit [ ] (16.8)
Remembering that 1Iris the bandwidth B of the receiver, we can rearrange
Eq. (16.8) as
nlT= B [Hz] (16.9 )
Thus the maximum number of resolution cells that can be searched every
second is set by the bandwidth of the system.
The simplest way to increase the bandwidth of a radar system is to use
multiple antenna beams and receivers; the effective bandwidth Beee is then
Beee = mB [Hz] (16.10)
where m = number of simultaneous beams and receivers.
322 UNDERSTANDING RADAR SYSTEMS
• Increasing the receiver bandwidth decreases the range resolution cell size
and cuts down the clutter.
• The use of more simultaneous beams implies that longer integration
times can be used in each beam position. This increases the radar
sensitivity.
• Longer integration times mean that the doppler resolution and clutter
rejection can be improved.
• The data processing rate for each target can be increased.
In this way, all our requirements can be met without increasing the
power x aperture product and exposing the radar to an increased risk of
ECM attack. Generally speaking, the analogue parts of a radar system are
not the factor limiting the bandwidth, but the computational power available.
This situation is changing with the arrival of parallel processing, very
large-scale integration (VLSI) signal processing integrated circuits and
optical methods of solving integrated circuit interconnection problems. Soon,
large amounts of low-cost computing power will be available, and we must
decide how to make best use of it in the design of the next generation of
radars. One of our main aims must be to make radars more adaptable to
the tasks they have to carry out.
16.3 ADAPTIVITY
In the past, radars have been rather unintelligent devices, perhaps more so
than was necessary. Early radars often did not know if they were being
THE FUTURE OF RADAR 323
jammed and yet the noise level in the receiver, measured continuously as
the antenna rotated, could have been used to draw a shape on a PPI display
representing the maximum detection range for targets of various Res. For a
radar that is internally noise-limited, these contours would be perfect circles
on a PPI display, but if external noise was received from any particular
direction, this would cause a dint in the circle and warn the operator that
the radar was not working well in that direction.
Imagine, now, a phased array radar in which all aspects of the system
are controlled by the data processing computer, as shown in Fig. 16.1. The
beam scheduling is controlled adaptively and well behaved targets are
inspected less frequently than those which have been found to be manoeuvring.
If the transmitted power is variable, this may also be controlled adaptively
to illuminate small long-range targets more strongly. It may be more
important to control the transmitter waveform adaptively and reserve long
pulse and wide swept bandwidths for difficult situations. When allocating
air defence resources, for example, it may be important to find out whether
a single echo represents one attacking aircraft or a very tight formation of
several planes; this may be possible by allocating extra integration time,
bandwidth and computing resources to the problem until it is resolved, while
other continuing searches are maintained using shorter pulses.
Adaptive doppler filtering and tracking algorithms may be one method
by which other information can be integrated into the system, in a process
known as data fusion. A modern warship, for example, has many sensors
physically close to each other and it is relatively easy (electrically) to
connect them together. If a target is identified by other means (e.g. visual
identification), the radar data processing can be allocated on the basis of
the known characteristics of the target. A modern fly-by-wire aircraft would
Figure 16.1 Adaptive phased array radar, in which all aspects of the system are controlled by
the data processing.
324 UNDERSTANDING RADAR SYSTEMS
ll~ Time/range
Figure 16.2 Searching for targets as a pattern among a series of scans can be more productive
than scanning each threshold independently. These scans were generated artificially using white
noise and a double pulse generator, set to give a low SNR. To find the targets, try looking at
the diagram with your eyes nearly level with the bottom of the page.
THE FUTURE OF RADAR 325
need more frequent inspection than a 852 bomber, because of the fighter's
greater manoeuvrability.
Data fusion is one of the areas where artificial intelligence (AI) techniques
can be used to make radars more intelligent. There is no magic in AI, but
there are some very useful programming ideas. The process of track initiation
could be improved by moving away from simple 'hit or miss' thresholding
to the utilization of' soft thresholding' and fuzzy logic processing. In addition,
pattern recognition algorithms could be used to determine whether a series
of poor-quality hits represents the presence of a real track or not. This process
is illustrated in Fig. 16.2; try looking at these range sweeps with your eye
almost level with the page and the two targets (both at very low SNR) should
become obvious. In fact, historical displays, which progressively reveal the
last few hits in a sector, have been used in the past, but it was the human
operator who made the decisions; AI will allow us either to give extra aid
to radar operators or to replace them by software, which is more reliable
and can detect targets at lower SNR for the same probability of detection
and false alarm as at present.
16.4 SUMMARY
The future of radar does not lie in larger and more powerful systems, but
rather in slightly smaller systems that are more agile, intelligent and difficult
to detect because of the larger bandwidths that will be used. The resolution
of radars, and the number of targets that can be tracked, can be expected
to increase as large amounts oflow-cost computer power become available.
The factor limiting radar performance is likely to remain technology
(rather than any law of physics) for some time to come. Until recently, it
was the ability to process the large amounts of data that created an upper
limit to performance, but with the advent of parallel processing the weakest
link in the chain is almost certain to become the A/D converter because of
the problem of increasing the sampling rate while maintaining a high dynamic
range. The recent trend of moving the point at which the A/D conversion
takes place further and further up the receiving chain towards the antenna
only exacerbates this problem.
Radar seems certain to provide challenging engineering, mathematical
and computational problems to be solved for years to come. We hope that
this book has conveyed the main ideas and helped you to understand the
underlying principles of radar. We hope also that you have gained an
appreciation of the importance of radar in many diverse areas, and sensed
some of the excitement of working in this field.
16.5 REFERENCE
I. Whither radar. M. F. Radford. GEe J. Res .. 3(2),137-143.1985.
APPENDIX
I
SYMBOLS, THEIR MEANING AND SI UNITS
These are the basic symbols used. Some have several subscripts, e.g. L, L.,
L.y " etc.
326
SYMBOLS, THEIR MEANING AND SI UNITS 327
II
ACRONYMS AND ABBREVIATIONS
AC alternating current
ACF autocorrelation function
A/D analogue-to-digital (converter)
AI artificial intelligence
ARM anti-radiation missile
ATC air traffic control
CAT clear-air turbulence
CCIR International Radio Consultative Committee
CFAR constant false-alarm rate
chirp linear frequency sweep during a pulse
COMINT communications intelligence
CW continuous wave
DC direct current
ECCM electronic counter-countermeasures
ECM electronic countermeasures
EJ escort jammer
ELINT electronic intelligence
EM electromagnetic
ERP effective radiated power
ERS-J European remote sensing satellite
ESM electronic support measures
EW electronic warfare
FFT fast Fourier transform
330
ACRo.NYMS AND ABBREVIATIo.NS 331
III
USEFUL CONVERSION FACTORS
333
334 UNDERSTANDING RADAR SYSTEMS
IV
USING DECIBELS
Decibels can be one of those little things that you can get a mental block
about. It is not that there is anything difficult about the maths, but rather
the concept of handling voltage and power ratios in this way. However, it
is worth learning to use decibels, because they make life so much easier for
the radar engineer.
The idea behind decibels is that the numbers used in radar calculations
are often very large (e.g. R4) or very small (e.g. Pr ), and they must usually
be multiplied and divided. This makes life difficult and can easily lead to
mistakes. Converting these numbers to logarithms means that, first, they
become a sensible size and, secondly, they only have to be added and
subtracted. This greatly facilitates repeating calculations, when one of the
parameters is varied.
The definition of decibels is that, if PI and P2 are two amounts of power
then the first is said to be n decibels greater than the second, where
n = IOlog lo (PdP2 )
The ratio of two voltages VI and V2 can be converted to decibels by
n = 2010glO(VI/V2)
In radar, it is more common to find the 10 loglo power version used in
calculations, rather than voltage ratios, but this is not necessarily so in other
subjects. To escape from decibels and return to linear numbers, divide n by
10 (or 20 for voltage ratios) and then use the IOXfunction on your calculator.
It can sometimes be difficult to understand negative decibels. These
represent ratios less than unity. A figure of - 10 dB means a ratio of 1/10,
335
336 UNDERSTANDING RADAR SYSTEMS
Ratio dB Ratio dB
(4rrj-3 -33.0 1 0
10- 3 -30.0 2 3.0
(4rr j-2 -22.0 4 6.0
10- 2 -20.0 5 7.0
10- 1 -10.0 10 10.0
1/4 -6.0 102 20.0
1/2 -3.0 103 30.0
1 MHz 60 dB Hz
1 GHz 90dB Hz
v
SOLUTIONS TO PROBLEMS
Chapter 1
1.1 (a) J./d = 0.03/(2.7 x 0.8) = 0.014 [rad] = 0.8°
(b) J.jd = 0.10/(3.6 x 0.8) = 0.035 [rad] = 2°
(Racal- Decca quote O.~o and 2° respectively in their specifications.)
337
338 UNDERSTANDING RADAR SYSTEMS
evaluation the error did not in fact exceed 6 m. Assuming a runway length
of 1 n. mile, this should be sufficiently accurate for a bad-weather landing.
1.7
P, 44 dBW 47.4 dB W
G,G, 64 dB 83.0dB
(J 10 dBm 2 10 dBm 2
•2
I. -30.5 dB m 2 -30.5 dB m 2
L, -5 dB -5 dB
Ij(4n)3 -33 dB -33 dB
IjN +131 dBW +131 dBW
Ij(SNR) -13 dB -13 dB
Chapter 2
2.1 Duty cycle = T/T = T x (PRF)
Mode 1: Duty cycle = 1.5 x 10- 6 x 480 = 1/1389
Mode 2: Duty cycle = 1.0 x 10- 6 x 691 = 1/1447
The mean power = peak power x duty cycle.
Mode 1: Mean power = 864 W
Mode 2: Mean power = 829 W
(Alenia quote 0.87 and 0.83 k W, respectively.)
2.2 We have
1.2 x 480 .
n = = 16 hIts/scan
6x6
2.3 Equating
R = C with
max 2 X (PRF) 12.4 X (PRF)
and remembering that 1 nautical mile = 1853 m, tells us that the number 12.4
is a nautical mile expressed in microseconds. This is another way of expressing
the '150 m per microsecond' rule as '12.4 J.lS for every nautical mile'.
2.4
PI = 60.8 dB W 1.2MW
GIGr = 73.0 dB G = 25000/1.2 X 4.7
(J = 3.0dB m 2 2 m 2 target
).2 = -12.7 dB m 2
L= -7.0dB
1/(4n)3 = -33.0dB
SNR = -10.0 dB
(Noise) -1 = + 145.8 dB W Assuming bandwidth = 1:- 1
= (1.5 X to- 6 )-1
R~ax = 219.9 dB m 4
Rmax = 55.0 dB m ==314km-170n.mile
The Alenia engineers' radar calculation worksheets use a more detailed
procedure, but arrive at a similar conclusion; 164 n. mile for this case.
8
2.5 c = 1.5 X to = 312.5 km
Runambig = 2 x (PRF) 480
This is similar to the maximum detection range and so the system is nominally
unambiguous in range. However, Alenia provides a 5 per cent PRF stagger
to remove long-range ground clutter effects caused by tropospheric ducting.
is not achieved because the sensitivity must be set such that the quantization
noise is small compared to the system noise.
(The ATCR-44K radar uses a 12 bit A/D converter sampling every
1.5 liS.)
2.7
Pt = 43.0dB W 20kW
GtGr = 32.0 dB
(1 = 3.0dB m 2
A. 2 = 7.6 dB m 2
L = -5.0dB
1/(4n)3 = -33.0dB
(SNR)-1 = -13.0dB
(Noise) -1 = + 147.0 dB W
181.6 dB m4
45.4dBm == 34.5 km == 18.6 n. mile
This is much lower than the modern Alenia radar.
To be fair to the Freya system, there would probably have been
considerable incoherent gain involved in the display process, which would
have improved the detection range over this single-pulse calculation. None-
theless, the calculation serves to illustrate that the improvement in the A. 2
factor over a modern L-band radar is not sufficient to compensate for the
much smaller power x aperture product.
Chapter 3
3.1 (PRF) x 60/(RPM) = 25 pulses/scan
3.2 The radar equation indicates a SNR of 0.3 dB for a 120 m 2 target at
100 km. In one minute, 1500 angular measurements are made, which would
increase the SNR of the measurement by up to 31.8 dB. The best that might
be achieved is 2° / j (2 x SNR) = 0.04°, although the spinning ofthe reflector
as it dangled below the balloon would cause RCS variations that would
degrade this figure. (Siemens- Plessey quote better than 0.15° for this system.)
3.5 (a) v,. (beam centre) = 300 cos 10° cos 45° = 208.9 m S-1
v,. (beam edge) = 300 cos 10° cos 47° = 201.5 m S-1
The difference of 7.4 m s - 1 corresponds to a doppler shift of ± 493 Hz across
the beam.
(b) 10 Hz = 0.15 m S-1 - 0.04°. These parameters are similar to those
used on the GEC-Ferranti Blue Vixen radar recently fitted to the UK Sea
Harrier aircraft. The Blue Vixen is a multimode fire control radar, but with
the capability of doppler beam sharpening. (GEC-Ferranti suggest that a
resolution of 1/20° or better is achievable with this technique.)
(c) v,. (ahead) = 300 cos 10° cos 0° = 295.4 m S-1
v,. (beam edge) = 300 cos 10° cos 2° = 295.3 m S-1
The difference in velocity is insufficient to give good resolution and doppler
beam sharpening is not normally used within ± 15° of the aircraft track.
Chapter 4
4.1 If mean RCS = s, PDF is
p(x) = (lIs) e-x!s x~0
Thresholding 10 dB above the mean gives
So expected number of false alarms is 512 2 X Pfa - 12. For gaussian PDF,
we must solve
t- cI>(k) = 4.54 x 10- 5
which gives k = 3.91.
In the exponential case we are nine (additive) standard deviations above
the mean; in the gaussian case only 3.91.
1/(5) V u(s - t)
O - : - - - - - t - - - r - - r - - - - - - - - - ----,
I
I
I
-T t- T T t+ T 5
By symmetry
It I > 2T
It I ~ 2T
(b) Energies are the same, therefore there is the same detection per-
formance after matched filtering. Output in case (i) is as above with T = 10- 3.
Output in case (ii), using diagrams as above, is:
SOLUTIONS TO PROBLEMS 343
Both outputs have the same maximum value at 0; this is the energy 2V2T.
The second waveform gives a sharper main lobe; this has implications for
system resolution, as discussed in Chapter 6.
4.4 1 MHz
14 ~I
A 1 1
1 1 H(w)
1 1
100 MHz
Use
Eh = - I fX [H(wW dw
210 - x,
Since the integral of each triangular function squared will not change if it is
moved to the origin, we can use the diagram below:
A
344 UNDERSTANDING RADAR SYSTEMS
2 x -1
2n
Ie
0
A2(l - OJ/C)2 dOJ =A-
C
3n
2
So
Output noise power = ~ x 106 X A2 N
If we use the second definition of bandwidth, we must use C = n x 106 ,
giving half the output noise power above.
4.6 In Chapter 1,
SNR = signal power
noise power
signal power
= x--------------------
B noise power / unit bandwidth
For simple pulses, 1/ B '" r: where r: = pulse duration. Therefore, first term
is signal energy.
f
t
r---------<r----~--~------~
[
1
I I
-t - T -t -T -t +T T
= exp( -jat 2 ) f T-
-T
t exp( -2jast) ds
exp( _jat 2 ) . •
. {exp[ -2Ja(T - t )t] - exp(2JaTt)}
2Jta
= ~ sin[(2T - t)at]
at
Pu(t) =
{ ~ sin[(2T -Itl)at]
at
o
It I :::; 2T
It I ~ 2T
This is plotted in the diagram.
346 UNDERSTANDING RADAR SYSTEMS
Pu (t)
T=1
a=5
-1
-2
Hence PI ~ 31.4 W.
SOLUTIONS TO PROBLEMS 347
Chapter 5
5.1 FFT gain is 15 dB plus a maximum incoherent averaging gain of 3 dB.
Manoeuvring not only alters the RCS, causing a loss of integration gain,
but also spreads the echo across several doppler cells, thereby making
detection difficult.
5.3 Equation (4.78) gives k = 4.8, i.e. a margin of 6.8 dB must be added to
the mean. Equation (4.81) gives a SNR of 19.9 or 13.0 dB.
5.4
Observation/time 2 3 4 5 6
5.5 Track error is 0.6 times measurement error after five observations. The
standard deviation is about 4 m, and Eq. (5.29) suggests an association gate
size of 6 m.
Chapter 6
6.1 (a)
V u(t)
p,,(t)
-TI2 TI2
Pu(t) = f"oo V 2
exp[ -(s + t)2jT2] exp( -s2jT2)ds
Using
0-
J I(21t) f oc'
-00
exp( -s2j20-2)ds = I
2
SOLUTIONS TO PROBLEMS 349
Td = 3T12
Output = X(t,O) = { °
I -ltllT It I ~ T
It I > T
Doppler frequency = 2n1T:
ejlttlT (lIn) sin(nltl/T) It I ~ T
Output = X(t, 2n1T) = {
° It I > T
In practice, the values in this question are not too realistic for aircraft. What
is the doppler velocity corresponding to T = I ms?
6.4 Using the notation of problem 4.2, and the answer to problem 6.1, gives
~!d = 4 T 13 (replace T by 2 T) for first pulse. For the second (see diagram)
350 UNDERSTANDING RADAR SYSTEMS
1.2
6.5 If there are M pulses, the energy both in u ( t) and in u' ( t) are M times
greater than for a single pulse. Hence using Eq. (6.3 ), the effective bandwidth
does not change.
u(t) = p(t) + p(t - T) + p(t - 2T) + p(t - 3T)
Therefore,
U(w) = (1 + e- jcoT + e- 2j(t)T + e- 3j(t)T)p(w)
1- e- 4j (t)T
= P(w)
1- e-j(t)T
-3;.,T!2 sin(2Tw) ( )
= e ' P w
sin(Tw/2)
Plots (shown below) of IP(w)1 for a = 1 and a = t, of Isin(2Tw )/sin(Tw/2)1
and of their product show that the statement does not hold for any of the
other commonly used definitions of bandwidth, e.g. 3 dB width or distance
to first zero in the transform. To make the statement reasonable, the pulses
2.0
lP(w)1 T= 'If
a= 1
-8 -6 6 8 w
352 UNDERSTANDING RADAR SYSTEMS
• 1'=0<;
4
3.S
3
2.S
I 2
1.S
O.S
onvoga
;T_a=l
i 3
0_
8 -6 -4 4 6 8
""'¥
T=7f
a = 1/2
IP(w)1
0.4
0.2
3.S
2.S
.,..'"'
-Ii
loS
O.S
omega
SOLUTIONS TO PROBLEMS 353
IP(w)1 = a; sa a:w)
2(
Chapter 7
7.1 The power received by a unity-gain antenna at 256 n. mile (~474 km)
would be '" -99 dB W. This is nearly 2 dB greater than the -101 dB
required to trigger the transponder, and the signal is therefore adequate
when the aircraft is in the centre of the beam. The excess power (1.7 dB)
354 UNDERSTANDING RADAR SYSTEMS
would ensure that, as the interrogation beam swept over the aircraft,
0
the transponder would be triggered for about 3/4 either side of maximum
illumination. This must be near the limit of satisfactory performance and is
probably the reason why Cossor quote 256 n. mile as the maximum range
of the system.
7.2 Pr = - 95.6 dB W at the centre ofthe beam. Over the azimuth beamwidth
of 2.45° the signal would equal, or exceed, -98.6 dB W.
7.5 From problem 7.2, Pr = 95.6 dB Wand SNR = 34.4 dB. We could there-
J
fore expect the intrinsic resolution of 2.45° to be improved by (2 x SNR)
to give an uncertainty of 0.03°. In practice, monopulse can improve upon
this figure when the aircraft is near the centre of the beam and the uncertainty
might be nearer half this value. A good guide to the dependence of the
angular error on the SNR, angle from bore-sight, etc., is to be found in
Chapter 10 of Stevens 1 •
Chapter 8
8.1 (a) 112.9km, (b) 130.4km, (c) 120.9km, (d) 233.2km.
8.3 (a) The detection would be halved to 2250 m by the 6 dB diffraction loss.
(b) The detection range of a point target would be reduced by a factor
of 0.71.
8.4 500 km. This principle is used in altimetry (see Chapter 11) to remove
ionospheric effects and allow precise measurements of ocean surface height
from space.
SOLUTIONS TO PROBLEMS 355
Chapter 9
9.1 Z = 200(1 )1.6 = 200 = 23[dB mm 6/m- 3 ], sO
Constant = - 106.4 [dB m s - 1] 2.3 X 1011 m S-1
Pt = 58.1 [dB W] 650kW
GtGr = 74 [dB] 37 dB each way
Z = 23 [dB mm 6 m- 3 ]
,'!.l) A¢ = -29.1 [dB rad 2] 2° beam
T = - 57.0 [dB s] 2 JlS pulse
1/).,2 = 19.7 [dB m- 2] 2.9GHz
1/ R2 = -86.0 [dB m- 2] 20 km range
Pr = -103.7 [dB W]
The SNR would therefore be 33.3 dB. More detailed performance calculations
by Met. Office engineers arrive at a similar figure of 32.6 dB.
The largest ratio is when r = 64 and R = 1 km, giving - 8.1 dB. The smallest
ratio is for r = 1/8 and R = 200 km, giving -97.5 dB. The dynamic range
requirement is therefore nearly 90 dB. This can be relaxed by using
range-dependent gain, and Siemens-Plessey do in fact use swept gain
following a 1/ R2 law to 200 km.
9.4 The R2 term increases the loss by only 0.1 dB and this is almost
compensated by an increase in the area of the layer illuminated by the
off-vertical beam. The cause of the 10 dB drop in power when viewing
off-vertical (an effect known as aspect sensitivity) may be an anisotropy of
the turbulence at the edges of the layer, although scattering from steps in
the refractive index may cause the same effect. Not all layers exhibit aspect
sensitivity.
9.5 We have
11 = 0.38C~)" -1/3
and so
C; = ARG'1
= 300 x 1259 x 2.1 x 10- 18 = 7.9 X 10- 13 = -12l.0dB
Therefore,
PI = 57.8 dB W
A2 = I5.0dB m 2
Lsys = -7.0 dB
C; = -121.0 dB
1/(16n2 ) = -22.0dB
I/R2 = -80.8dBm- 2
Pr = -158.0dB W
9.6 The MU transmitter power is 2.2 dB larger than for the SOUSY system.
The antenna gain, at 34 dB, is 3 dB larger, and so the SNR should be about
5.2 dB larger. This illustrates the efficiency of performing calculations in dB,
rather than multiplying linear terms.
9.7 The PRF must be 6250 Hz and the total number of pulses transmitted
in 10.5 s is 65536 (see also Eq. (10.1) in the next chapter). Averaging these
echo pulses in blocks of 512 leaves 128 samples for the FFT processing. Each
FFT cell will be 1/10.5 Hz wide and, as there are 128 points, the doppler
spectrum will be ± 6 Hz wide.
9.9 Yes. The calculations in Section 9.4 hold for MST radar and show
meteors to be very bright targets, although somewhat short-lived compared
to the integration times usually used by MST radars. For a meteor to be
observed in a narrow vertical beam of an MST radar, it must be travelling
horizontally with a considerable path length through the atmosphere. Under
these cond!tions, many small meteors burn up before they reach the radar
beam and some larger meteors disintegrate through thermal shock. For these
reasons, although MST radars are used to observe meteors, few meteor
echoes are observed in the vertical beam.
Chapter to
to.t SRI quote the following figures: (a) 0.5°, (b) gain exceeds 30 dB between
8 and 28 MHz, (c) 93 dB W, (d) FOM of 104 dB J.
SOLUTIONS TO PROBLEMS 357
10.4
4n = 11.0 dB
PtGtGrt; = 104 dB J
(1 = 52.9 dB m 2
m= 6 dB
Is = -15 dB
l/Fa = -30 dB
l/kTo = +204 dB W- 1 Hz
1/UhA)2 = -275 dB m-2
SNR= 57.9 dB
Chapter 11
11.1 The dominant Bragg line should appear at ±0.4 Hz in the doppler
power spectrum. The extra shift of 0.3 Hz implies a tidal current of 3 m s - 1
or nearly 6 knots. Ship speeds are generally restricted to 5 knots approaching
a harbour, and hence the knowledge of a 6 knot cross-current could aid the
pilot.
2
-[(R+rs)-R]
c
Hence
2rs ' '-- ,.'
"
c B it.
358 UNDERSTANDING RADAR SYSTEMS
The range difference between near and far edges of the beam (in slant range) is
h h
RM-R = ---
m cos(O + P) cos (J
-h( cos 0 -
1P __1)
sin 0 cos 0
hP sin (J
cos 2 0
Therefore
2B
N - - hP tan 0 sec 0
c
Using values B = 5 X 108 , P= 0.06/0.6 = 0.1 rad, then
N-8secOtanO
The variation of N with 0 is shown in the plot below.
I:~
20t
~ 1;L-~---1-~~--L--....,.L---~
Incidence angle [deg]
~~p2
c 8
Hence condition given.
For ERS-l: p = )./d = O.ot8 rad or 1.04 and R 0
= 7.85 x lOs m. Thus,
for pulse-limited operation,
r ~ 1.07 x 10- 7 s
The radar equation is
P'G 2C1A).2L
P, = I s
r (41t)3R 4
whereC1 = ReS/unit area, A = illuminated area, G = 41tA;/).2 and r = antenna
radius. So
P,1t 2r4 C1cr).2L
P, = I s
r 2R3
For a simple pulse and matched receiver, noise power = kToB - kTo/t, giving
212L
SNR = P,In: r C1cr.ll. s
2 4
3
2R kTo
= 0.0375PI C1
= 59.3 (17.7 dB)
for PI = 500 Wand C1 = 5 dB m 2 • In practice, the ERS-l altimeter uses a
20 ps chirp with 50 W Tx power.
Ma = Ra v" (d s )3
Ms Rs V. da
1 1
=-x-xlO00
20 135
- 0.37
At C-band, Ma = t X 106
At L-band, Ma = i X 106
11.5 The return from the slant swath will occupy a time duration
2 2Rs
-(RM - Rm)=-
c c
Successive pulses must be at least this far apart, even with no margin, i.e.
PRF ~ c/2R s
But PRF ~ 2V/d, and so
Rs ~ cd/4V
For the aircraft:
8
R & 3 X 10 X 2 x 3600 = 180 km
s -..::: 4 x 300 x 1000
In practice, aircraft SARs normally use much narrower swaths. Air-borne
SARs therefore do not usually have to trade swath width for resolution.
By contrast, for the spacecraft:
3 5
d >- 4 x 7.5 X 10 x 10 = 10 m
r 3 X 108
Hence best possible resolution is 5 m in the along-track direction. The
trade-off between coverage and resolution is a serious design consideration.
11.6 Lay-over occurs when the top of a slope is nearer to the radar than
the bottom (see diagram). The limiting case occurs when the top and bottom
(and the whole of the slope) are imaged in the same place, i.e., are equidistant
from the radar. This is the line OX in the diagram. At the long ranges used
by SAR, this means that OS is nearly perpendicular to OX, or <p = (). Hence
any slope facing the satellite with a gradient exceeding 23° will suffer lay-over.
Slopes of this magnitude are common in hilly areas.
SOLUTIONS TO PROBLEMS 361
For the aircraft, incidence angles range from 78.7° to 84.3°. Hence
lay-over will occur only in very steep terrain, but shadowing will be common.
SAR operating at these angles can respond to very slight variations in
topography, such as you see near sunset as shadows lengthen. Note that for
the resolutions of the order of 1 m or so available to air-borne radar, lay-over
is always encountered in images of urban areas.
Chapter 12
12.1 200 V /10 mV implies a dynamic range of 86 dB, reduced to 69 dB by
the antenna losses. The attenuation rate and the depth requirement suggest
a 99 dB dynamic range. The signal must therefore be boosted by at least
30 dB if targets are to be detected to depths of 4.3 m. These figures are, in
fact, those given in reference 6, in which it is reported that objects were
detected to 3.5 m (the limiting factor being mainly television interference).
12.2 11.6 MHz (PRFs this high are not used in practice).
362 UNDERSTANDING RADAR SYSTEMS
12.3 The signal appearing in the receiver may be only 17 dB below 200 V, or
as high as 28 V (reference 6). This would be sufficient to damage the receiver
if protection methods (attenuators and diode limiters) were not used.
Chapter 13
13.1 (a) 1 kHz.
(b) 2 kHz.
(c) Monostatically the target motion decreases the path of the radio
signal on both the outward and the return leg. In this bistatic case, only the
outward path is shortening.
13.2 (a) 60 km (remember to allow for the time taken by the direct signal
travelling along the baseline).
(b) 39.7 km.
(c) 20.3 km.
13.4 (a) 6.7 km (about 22000 ft). In practice, some pulse integration would
be needed for the doppler processing and the extra gain from this would
enable the aircraft to loiter at much higher altitudes.
(b) The direct signal would be about 41 dB higher than the echo. This
should cause no dynamic range problems for the receiver.
(c) A radar on board the aircraft would be able to use the 20 dB
antenna gain both ways. This would more than compensate for the additional
path losses, allowing a smaller transmitter to be used (assuming the same
RCS). However, it would give away the aircraft position.
Chapter 14
14.1 10.5 dB.
14.2 40 dB to -104 dB W.
SOLUTIONS TO PROBLEMS 363
14.370mW.
14.6 A suitable PRF might be 100kHz (unambiguous range 1.5 km). The
answer to problem 14.5 implies an integration time of 0.1 s at this PRF. This
gives a speed resolution of 0.15 m s - 1, or 0.3 mph, sufficient to determine
whether a car is speeding.
14.7 (a) Radar warning receivers (RWRs) covering several bands used by
speed traps can be purchased, but they are expensive.
(b) A jamming device would be illegal and very expensive, unless the
exact radar frequency was known.
(c) Stealth materials are also expensive and only really effective against
a radar of known frequency. In any case, it would be difficult to conceal
some parts of the car such as the headlights, driver, radio antenna, etc.
Perhaps the cheapest solution is to keep to the speed limit. This also sa ves fuel.
Chapter 15
15.1 The antenna gain can be derived from either 475 (26.8 dB) plus 7.24 dB
for each element, or from G = 4nAI A2 , using the array area of 8330 m2 • Both
methods give a gain of 34 dB. The beamwidth AID = 6.45/103 = 3.6° is the
value quoted in the article.
15.3 Roughly 56° off-zenith for the vertical beam. The authors quote 'no
grating lobe is formed at beam positions within 40° of the zenith '.
15.4 Calculating 475 phase angles and setting up 475 phaseshifters 2500
times per second would require over one million such operations per second.
Instead, the MU system stores all phaseshift data required for beam steering
in a ROM (read-only memory) installed in each transmit module. The
appropriate values are then read out according to instructions from the
module controller.
BIBLIOGRAPHY
Throughout this book we have tried to restrict references to books which are relatively
easy to obtain. Occasionally we have referred to original publications in journals,
but these can be found in most university and industrial libraries. To help with finding
references, we list them for you (first author only) alphabetically below.
364
BIBLIOGRAPHY 365
368
INDEX 369