atmospheric propagation

215
ATMOSPHERIC PROPAGATION EFFECTS ON HETERODYNE-RECEPTION OPTICAL RADARS by DAVID MICHAEL PAPURT B.S.E.E., B.A., University (1977) S.M., Massachusetts Institute (1979) E.E., Massachusetts Institute (1980) SUBMITTED OF THE of Toledo of Technology of Technology IN PARTIAL FULFILLMENT REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY May 1982 Massachusetts Institute of of Author.., Department of Electrical Certified by........'......... Technology, 1982 Engineering/and Computer Science May 14, 1982 Jeffrey H. Shapiro Tjhsis Supervisor Accepted b.......... Arthur C. Smith Chairman, Department Committee on Graduate Students Archives MA SSACHUSETTS NSitTZTC OF TECHNOLOGY OCT 20 1982 UBRARIES 0 Signature

Upload: others

Post on 25-Feb-2022

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: ATMOSPHERIC PROPAGATION

ATMOSPHERIC PROPAGATION EFFECTS

ON

HETERODYNE-RECEPTION OPTICAL RADARS

by

DAVID MICHAEL PAPURT

B.S.E.E., B.A., University(1977)

S.M., Massachusetts Institute(1979)

E.E., Massachusetts Institute(1980)

SUBMITTEDOF THE

of Toledo

of Technology

of Technology

IN PARTIAL FULFILLMENTREQUIREMENTS FOR THEDEGREE OF

DOCTOR OF PHILOSOPHY

at the

MASSACHUSETTS INSTITUTE OF TECHNOLOGY

May 1982

Massachusetts Institute of

of Author..,Department of Electrical

Certified by........'.........

Technology, 1982

Engineering/and Computer ScienceMay 14, 1982

Jeffrey H. ShapiroTjhsis Supervisor

Accepted b..........Arthur C. Smith

Chairman, Department Committee on Graduate Students

ArchivesMA SSACHUSETTS NSitTZTC

OF TECHNOLOGY

OCT 20 1982

UBRARIES

0

Signature

Page 2: ATMOSPHERIC PROPAGATION

-2-

ATMOSPHERIC PROPAGATION EFFECTS

ON

HETERODYNE-RECEPTION OPTICAL RADARS

by

DAVID MICHAEL PAPURT

Submitted to the Department of Electrical Engineering & Computer Scienceon May 14, 1982 in partial fulfillment of the requirements for the

Degree of Doctor of Philosophy

ABSTRACT

The development of laser technology offers new alternativesfor the problems of target detection and imaging. The performance ofsuch systems, when operated through the earth's atmosphere, may beseverely limited by the stochastic nature of atmospheric opticalpropagation; that is, by turbulence, scattering, and absorption. Amathematical system model for a compact heterodyne-reception laserradar which incorporates the statistical effects of target speckleand glint, local oscillator shot noise, propagation through eitherturbulent or turbid atmospheric conditions, and beam wander ispresented. Using this model, results are developed for the imagesignal-to-noise ratio and target resolution capability of the radar.Complete statistical characterizations for the radar return aregiven. An experiment and data processing techniques, aimed atverifying the above statistical models, are described and resultsgiven. Notable here is the verification of the lognormal characterof the turbulent atmosphere induced fluctuations on the radar return.Target detection performance of the radar is investigated. Regimesof validity for the various target return models are established.

Thesis Supervisor: Jeffrey H. Shapiro

Title: Associate Professor of Electrical Engineering

Page 3: ATMOSPHERIC PROPAGATION

To my MotheA and FatheA

-ot the%'L eoAnt,6, patience, and Zove

Page 4: ATMOSPHERIC PROPAGATION

-4-

ACKNOWLEDGEMENTS

It has indeed been an honor and a pleasure to be associated

with my thesis supervisor, Professor J.H. Shapiro. His guidance

throughout the course of my stay at M.I.T. has been invaluable. I

would also like to thank my thesis readers Dr. R.C. Harney and

Professor W.B. Davenport. Their many suggestions have improved this

work significantly. Also, W.B. Davenport, in his role as my

graduate counselor, provided advice, without which, this goal would

never have been realized.

I would like to acknowledge all members of the Optical

Propagation and Communication research group at M.I.T. The many

stimulating discussions, some of a technical nature and many

nontechnical, between myself and members of this group have added

to this work. In particular, I would like to mention Dr. J. Nakai.

I feel fortunate to have made his acquaintance and honored to be his

friend.

Many members of the Opto-Radar Systems group at M.I.T.

Lincoln Laboratory played a role in this research. R.J. Hull,

T.M. Quist and R.J. Keyes deserve special mention for their efforts

in providing me with radar data and the computer facilities to process

this data with.

Page 5: ATMOSPHERIC PROPAGATION

-5-

Financial support for my doctoral studies has been provided

by the U.S. Army Research Office, Contract DAAG29-80-K-0022. This

support is gratefully acknowledged.

Deborah Lauricella has my appreciation for her pains in

typing this document.

Page 6: ATMOSPHERIC PROPAGATION

-6-

TABLE OF CONTENTS

Page

ABSTRACT......

ACKNOWLEDGEMEN

LIST OF FIGURE

LIST OF TABLES

CHAPTER I.

CHAPTER II.

CHAPTER III.

CHAPTER IV.

..................................

TS ..........................

S .... .... ................ ......

............................. 0.....

INTRODUCTION.....................

A. Laser Radar Configuration....

ATMOSPHERIC PROPAGATION MODELS...

A. Free Space Model............

B. Turbulence Model.............

C. Turbid Atmosphere Model......

D. Backscatter..................

TARGET INTERACTION MODEL.........

A. Planar Reflection Model......

B. Relationship to Bidirectional

SCANNING-IMAGING RADAR ANALYSIS..

A. Single Pulse SNR.............

B. Speckle Target Resolution inVisibility Atmosphere........

C. Identification of Atmospheric

D. Turbulence SNR Results.......

E. Low Visibility SNR Results...

Reflectance....

................

................

the Low................

Effects ........

................

................

F. Beam Wander Effects..........................

G. Correlation of Simultaneous Speckle TargetReturns......................................

H. Backscatter..................................

2

4

8

14

15

17

27

30

31

32

34

35

35

37

40

41

45

50

52

56

71

87

90

. .. ... . .. .. .. . ..

. . ... .. .. .. . .... 0

Page 7: ATMOSPHERIC PROPAGATION

-7-

CHAPTER V.

CHAPTER VI.

CHAPTER VII.

CHAPTER VIII

APPENDIX A.

THEORY VERIFICATION............................

A. Laser Radar Description....................

B. Data Analysis Techniques...................

C. Beam Wander................................

0. Turbulence.................................

TARGET DETECTION...............................

A. Problem Formulation........................

B. Single Pulse Performance...................

C. Multipulse Integration.....................

D. Multipulse Performance.....................

SYSTEM EXAMPLES................................

A. EXAMPLES...................................

SUMMARY........................................

DERIVATION OF THE MUTUAL COHERENCE FUNCTION(MCF)..........................................

REFERENCES....................................................

BIOGRAPHICAL NOTE.............................................

Page

96

96

99

102

113

135135

138

147

150

165

174

194

197

211

215

Page 8: ATMOSPHERIC PROPAGATION

-8-

LIST OF FIGURES

Figure Page

1.1 Laser radar configuration.............................. 18

1.2 Transmitted power waveform............................. 21

1.3 Heterodyne receiver model.............................. 25

2.1 Turbulence field coherence length p0 vs. propagation

path length L for conditions of weak turbulence,

moderate turbulence and strong turbulence.............. 28

3.1 Geometry for defining bidirectional reflectance

p'(x;Tf,T ) '..' '' ' 39

4.1 Reflected spatial modes from resolved and unresolved

speckle targets........................................ 61

4.2 Reflected phasefronts from glint targets............... 64

4.3 Target return PDF for beam wander fluctuations,

Gaussian beam, uniform circular beam center,

|m/rb = 0.0, R/rb = 1.0 ........................... 77

4.4 Saturation SNR due to two types of beam wander

fluctuation............................................ 78

4.5 Target return PDF for beam wander fluctuations;

fan beam, uniformly distributed beam center,

m/rb = .45, R/rb = 1.2................................. 82

4.6 Saturation SNR due to fan beam, uniform beam center

fluctuations........................................... 83

4.7 Target return PDF for beam wander fluctuations;

fan beam, Gaussian distributed beam center, m/rb = 0.8,

a /rb = 1.25........................................... 85

Page 9: ATMOSPHERIC PROPAGATION

-9-

LIST OF FIGURES

Figure

4.8 Single scattering layer..................................... 91

5.1 Normalized histogram of 100 consecutive retro returns

taken in full scanning mode and theoretical PDF

Eq. (4.F.23), R/rb = 1.3, m/rb = 0.0........................ 105

5.2 Normalized histogram of 200 consecutive returns taken in

full scanning mode and theoretical PDF Eq. (4.F.23),

R/rb = 1.2, m/rb = 0.5...................................... 106

5.3 Normalized histogram of 200 consecutive retro returns taken

in full scanning mode and theoretical PDF (4.F.23),

R/rb = 1.3, m/rb = 0.56..................................... 107

5.4 Normalized histogram of 200 consecutive retro returns from

the "side" pixel of Figure 5.2 taken in full scanning mode

and theoretical PDF (4.F.28), R/rb = 1.2, m/rb = 1.43....... 110

5.5 Normalized histogram of 400 consecutive retro returns

taken in full scanning mode and theoretical PDF (4.F.15),

R/rb = 1.4, m/rb = 0.6...................................... 111

5.6 Normalized histogram of 100 "hot spot" retro returns

taken in reduced scanning mode and the lognormal PDF

(5.D.l), a2 = .0045......................................... 114x

5.7 Time evolution; Row location of hot spot in the l00-60x128

pixel frames used in the example of Figure 5.6.............. 116

5.8 Time evolution; Column location of hot spot in the

l00-60x128 pixel frames used in the example of Figure 5.6... 117

5.9 Normalized histograms of 300 "hot spot" retro returns

taken in reduced scanning mode and the lognormal PDF

(5.D.1), a2 = 0.0138........................................ 119x

Page

Page 10: ATMOSPHERIC PROPAGATION

-10-

LIST OF FIGURES

Figure Pag

5.10 Normalized histogram of 400 "hot spots" retro returns

taken in reduced scanning mode and the lognormal PDF

(5.D.1), &2 = 0.018...................................... 120x5.11 Normalized histogram of 300 "hot spot" polished sphere

returns taken in reduced scanning mode and the lognormal

PDF (5.D.1), g 2 = .0083.................................. 121x

5.12 Normalized histogram of 300 "hot spot" polished sphere

returns taken in reduced scannina mode and the lognormal

PDF (5.D.1), a'2 = 0.004.................................. 122x

5.13 The theoretical curve (4.D.6) and estimates SNRSAT from

the data of Figures 5.6,5.9-5.12......................... 124

5.14 Normalized histogram of 2000 squared, consecutive

speckle plate returns taken in full scanning mode and

exponential PDF.......................................... 125

5.15 Normalized histogram of 400 consecutive speckle plate

returns taken in full scanning mode and Rayleigh PDF..... 126

5.16 Normalized histooram of 1200 speckle plate returns

taken in reduced scanning mode and Rayleigh PDF.......... 128

5.17 Normalized histograms of 1400 consecutive speckle plate

returns taken in staring mode and Rayleigh PDF........... 129

5.18 Normalized histogram of 300 speckle plate returns taken

in full scanning mode and Rayleigh PDF...................131

5.19 Normalized histogram of the same 300 speckle target

returns as Figure 5.18 and a Rayleigh times lognormal

PDF, a 2 = 0.01........................................... 132

Page 11: ATMOSPHERIC PROPAGATION

-11-

LIST OF FIGURES

Page

6.1 The probability density function, p (Y) = 2K (2VY).......142

6.2 Single pulse detection probability vs. CNR for a glint

target in bad weather.....................................144

6.3 Single pulse receiver operating characteristics for a

glint target in bad weather...............................145

6.4 Single pulse detection probability vs. CNR for a single

glint target in free-space, three levels of turbulent

fluctuations and bad weather. P F = 10- throuhout.......146

6.5 Likelihood ratio R (Eq. (6.C.3)) and parabola

log R = 3.1 + 0.3441r' vs, matched filter envelope

detector output Ir.......................................149

6.6 Threshold y vs. number of pulses necessary to maintain

-= 102 154PF

6.7 Single-pulse and multipulse detection probability vs. CNR

for a glint target in bad weather. PF = 104 throughout.. 158

6.8 Single-pulse and multipulse detection probability vs.

CNR for a glint target in bad weather. PF=10-12 throughout. 159

6.9 The number of pulses M necessary to achieve PF = 10-12

pD= .99 vs. CNR for a glint target in free-space, two

turbulent fluctuation levels, and bad weather.............160

6.10 The number of pulses M necessary to achieve PD = .99 and

two different false alarm probabilities vs. CNR for a

glint target in low visibility............................161

6.11 Ten pulse detection probability vs. CNR for a glint

target in several turbulent atmospheres and a scattering

atmosphere. PF = 10l2 throughout........................163

Page 12: ATMOSPHERIC PROPAGATION

-12-

LIST OF FIGURES

Figure Paae

6.12 Fifteen pulse detection probility vs. CNR for a glint

target in several turbulent atmospheres and a scattering

atmosphere. PF = 10-12 throughout........................164

7.1 Geometry for a single scattering layer between radar

and target................................................170

7.2 Geometry for a scattering layer near the target...........171

7.3 Normalized backscatter power from a uniform scattering

profile vs. t................... ......................... 175

7.4 Maximum normalized backscatter power from a scattering

layer L meters from the radar............................177

7.5 Extinguished free-space and MFS resolved speckle target

CNR vs. target range for the CO2 system and a uniform

scattering profile........................................178

7.6 Extinguished free-space and MFS resolved speckle target

CNR vs. tarqet range for the Nd:YAG system and a uniform

scattering profile........................................179

7.7 Atmospheric beamwidth for a single layer scattering

profile vs. layer thickness...............................181

7.8 Extinguished free-space and MFS resolved speckle target

CNR for the CO2 system and a single scattering layer vs.

layer thickness...........................................182

7.9 Extinguished free-space and MFS resolved speckle target

CNR for the Nd:YAG system and a single scattering layer

vs. layer thickness......................................183

7.10 Atmospheric beamwidth for a scattering layer near the

target and the Nd:YAG system vs. layer thickness......... 186

Page 13: ATMOSPHERIC PROPAGATION

-13-

LIST OF FIGURES

Figure Page

7.11 Extinguished free-space and MFS resolved speckle target

CNR for the Nd:YAG system and a scattering layer near the

target vs. layer thickness............................... 187

7.12 Field coherence length for a scattering layer near the

target and the Nd:YAG system vs. layer thickness......... 190

7.13 Extinguished free-space and MFS unresolved glint target

CNR for the Nd:YAG system and a scattering layer near

the taroet vs. layer thickness........................... 191

A.1 Geometry relating to the definition of specific

intensity................... ............................ 198

A.2 Geometry for relating the specific intensity and the

mutual coherence function................................ 201

A.3 MCF's for plane wave input............................... 207

A.4 Real phase function...................................... 209

Page 14: ATMOSPHERIC PROPAGATION

-14-

LIST OF TABLES

Table

7.1

7.2

7.3

7.4

CO2 Laser Radar System Parameters...................... 167

Nd:YAG Laser Radar System Parameters................... 168

Atmospheric Parameters at CO2 Laser Wavelength......... 172

Atmospheric Parameters at Nd:YAG Laser Wavelength.

Page

..... 173

Page 15: ATMOSPHERIC PROPAGATION

-15-

CHAPTER I

INTRODUCTION

The development of laser technology offers new alternatives to

the problems of target detection and imaging. Among the advantages

provided by laser radars over conventional radar systems are increased

angular, range, and velocity resolution with compact equipment. However,

performance of the laser system may be severely limited by the stochastic

nature of atmospheric optical propagation; that is, by turbulence,

absorption, and scattering. Herein, a mathematical model describing a

heterodyne reception optical radar is presented. The model incorporates

the statistical effects of propagation through tubulent and turbid

atmospheric conditions, as well as target speckle and glint, and local

oscillator shot noise.

A convenient theoretical model describing optical propagation

through atmospheric turbulence has been established [1]. In contrast,

optical propagation through bad weather is much more difficult to model

and to date no comprehensive theory exists which characterizes this

propagation regime in complete generality. From the available turbid

atmosphere propagation models the multiple forward scatter (MFS) extended

Huygens-Fresnel formulation [2] is useful in our application. Both the

turbulence model and MFS model are expressed in linear-system form in

which the random nature of the propagation process is represented by a

Page 16: ATMOSPHERIC PROPAGATION

-16-

stochastic point source response function (Green's function). Such a

linear model leads to a tractable overall radar system model.

In this thesis, we consider the performance of a scanning imaging

system and to a lesser degree that of a target-detection system. By

means of the turbulence model, the primary atmospheric effect limiting

compact* radar system performance has previously been shown to be

scintillation [3-5]. In contrast, beam spread and receiver coherence

loss as well as scintillation will be shown to be important when the MFS

model applies.

To validate the theory measurements made on the compact C02-laser

radar, developed as part of the M.I.T. Lincoln Laboratory Infrared

Airborne Radar (IRAR) project, have been made available. The compact

laser radar system [6] employs a one-dimensional, twelve-element HgCdTe

heterodyne detector, a transmit/receive telescope of 13 cm aperture, and

a 10 W CO2, 10.6 vm laser, which is operated in pulsed mode. For each

pulse, the intermediate frequency (IF) portion of the heterodyne detector

outputs are digitally peak-detected to yield 8 bit range and intensity

values. In the case of a large signal return the intensity value is

essentially the output of a matched filter envelope detector. Hence,

these data can be compared to the theory.

An outline of the topics covered herein is as follows. First, in

this chapter, a description of the laser-radar configuration will be

presented. This will be followed by descriptions of the atmospheric

*The term "compact" indicates a system that can be installed on a vehiclesuch as a truck or airplane.

Page 17: ATMOSPHERIC PROPAGATION

-17-

propagation and target interaction models, which appear in Chapters II

and III, respectively. The performance of a scanning-imaging radar in

both turbulence and low visibility will be considered in Chapter IV. As

speckle target returns from disjoint diffraction-limited fields-of-view

(FOV) will be shown to be uncorrelated in low-visibility weather, and

each picture element (pixel) of the image is assumed to encompass at

least a diffraction-limited FOV, we need consider only single-pixel

performance. This will be achieved via a signal-to-noise ratio (SNR)

analysis. Also, speckle target resolution in bad weather is considered

here. In Chapter V, theory verification efforts are described.

Single-pulse and multi-pulse target detection is explored in Chapter VI

with emphasis given to low visibility results. In Chapter VII system

examples are presented. Finally, in Chapter VIII, we summarize our

results and present suggestions for future work.

A. Laser Radar Configuration

A model of the laser radar configuration is shown in Figure 1.1.

A series of laser pulses ET(Pt) propagating nominally in the +z

direction are transmitted from the radar located in the z = 0 plane, and

illuminate a target located in the z = L plane (Figure 1.la). A fraction

of the illuminating field Et(PI',t) is reflected in the -z direction back

towards the radar (Figure 1.1b). The nature of this reflected field

E r(',t) clearly depends upon the reflection characteristics of the target,

as does the received field ER(p,t). This recieved field is mixed with a

local-oscillator field E,(P,t) and focussed onto an optical detector. The

Page 18: ATMOSPHERIC PROPAGATION

W mw

TRANSMITTER

TRANSMITTEDBEAM

E T(-P, t) a) FORWARD PATH

ILLUMINATINGBEAM

Et(-p, t)

0

ATMOSPHERE

RECEIVEDBEAM

ER(p ,t)b) RETURN PATH

REFLECTEDBEAM

Er(pt)

TARGET

z= L

TARGET

Figure 1.1: Laser Radar Configuration

L

RECEIVER

lw w

A TMOSPHERE D

Page 19: ATMOSPHERIC PROPAGATION

-19-

intermediate frequency (IF) component of the photocurrent comprises

the observed signal.

To as large an extent as possible the analysis has been tailored

to the IRAR project's compact C02-laser radar system. Thus, the

transmitter and receiver are taken to be co-located with common

entrance/exit optics of aperture diameter from 5 to 20 cm. The transmitter

will be assumed to produce a periodic train of rectangular-envelope laser

pulses while the local oscillator operates cw producing an ideal

monochromatic wave offset in frequency by the intermediate frequency

v IF* Targets are assumed to lie along line-of-sight paths a distance L

from the radar where 1 km < L < 10 km. In accordance with the above

conditions and transmitter pulse durations and pulse repetition frequencies

anticipated in real radar scenarios [7,8] we have the following

characterization.

1. The transmitted field has a quasi-monochromatic, linearly

polarized electric field proportional to

ET(Pt) = Re{u (p,t) e 0 } (l.A.1)T

where v is the optical carrier frequency, and P = (x,y) is

a position vector transverse to the direction of propagation.

The complex envelope uT(P,t) is expressed as the product of

a normalized spatial mode F(P) and a time waveform sT(t)

whose magnitude square is the transmitted power PT(t) (Watts).

Page 20: ATMOSPHERIC PROPAGATION

-20-

uT(P,t) = FT() -T(t) (1.A.2)

PT(t) = Is-T(t)12 (l.A.3)

dPIT( 12= 1 (1 .A.4)

This implies that 1u-T(P,t)I2 is the transmitted power density

(Watts/m 2). The transmitted power waveform PT(t) is assumed

to be as shown in Figure 1.2 where t is the pulse duration

and l/T the pulse repetition frequency.

2. We assume a scalar wave theory to describe the propagation.

We also assume the pulse duration to be short in comparison

to an atmospheric coherence time Tc >> tp and long in

comparison to the reciprocal coherence bandwidth (multipath

spread) of the atmosphere 1/Bcoh < t . The complex envelope

of the illuminating field in the z = L plane can then be

represented by the linear superposition integral

ut(P',t) = dp hLF(PiP) T(p,t - L/c) (l.A.5)

where hLF(P'P) is the stochastic atmospheric Green's function

(point source response) and c is the propagation velocity of

light. The subscripts indicate a path length of L meters in

the +z (or forward) direction. In free space (l.A.5)

becomes the Huygens-Fresnel integral [9]. In the atmosphere

Page 21: ATMOSPHERIC PROPAGATION

-21-

PT (t)

0 t T p +t 2 , 2 c +Pp p p p p p

Figure 1.2: Transmitted Power Waveform

Page 22: ATMOSPHERIC PROPAGATION

-22-

hLF(PI',P) is random with statistical characterization

dependent upon weather conditions.

3. A planar target-interaction model is assumed so that for a

stationary target the complex envelope of the reflected field

is

I-r (PI',t) = ut (_',t) T(-') (l.A.6)

where T(P') is the field reflection coefficient at the

point P'. In general T(p') is a random function containing

specular and diffuse components.

4. As before, we can represent the received field as a

superposition integral

u (_gPqt) = d' h LR , ) -r(p',t - L/c) (l.A.7)

In (l.A.7) hLR(P,P') is aoain the stochastic atmospheric

Green's function where the subscript R refers to propagation

in the -z (or return) direction. In turbulence, and in low

visibility when the target range is sufficiently short, the

propagation medium is reciprocal [10,11], i.e.,

hLF(PP') = hLR(', ((1.A.8)

Page 23: ATMOSPHERIC PROPAGATION

-23-

In low visibility and with a suitably large target range

we take hLF and hLR to be statistically independent. This

is due to the atmosphere decorrelating between pulse

transmission and reception times, and is an appropriate

assumption when the atmospheric coherence time Tc is shorter

than the roundtrip propagation delay 2L/c. We will refer

to the former case as "effectively monostatic" and the latter

case as "effectively bistatic" regardless of the radar

configuration.

5. According to the antenna theorem for heterodyne reception

[12], we can describe the photodetection process as though

it were taking place in the receiver's entrance pupil. Thus,

the detected field is taken to be ER(Pt) + EZ(p,t) where

the cw field E, has complex envelope

j27rv tu(-,t) = P1 F (P) e IF (l.A.9)

In (l.A.9), vIF is the intermediate frequency, PZ the local

oscillator average power and F a normalized spatial mode.

The photocurrent is passed through a rectangular passband

filter of bandwidth 2W centered at vIF. It is straightforward

to show that under the strong local oscillator condition [13]

a normalized (i.e. proportional to the photocurrent) IF

signal r(t) can be expressed as filtered signal plus noise.

Page 24: ATMOSPHERIC PROPAGATION

-24-

This is shown in Figure 1.3 where the pass band filter H(f)

has the input with IF complex envelope

y(t) = d -R(t) F*(P) + n(t) (l.A.10)~hvJ 0 -

The additive noise n(t) is a zero-mean, white, circulo-complex

Gaussian process with

<n(t) n*(s)> = tp 6(t - s) (l.A.ll)

where <-> denotes ensemble average, hv0 is the photon energy

and q is the detector quantum efficiency. Substituting for

u_,t) in (l.A.10) and making the definitions

EF(I') d- h LF(PI',P) F T(P) (l.A.12)

P' = d h R ,1 )' *( ) ( .A.13)

the IF complex envelope (l.A.10) becomes

y(t) = 2 sT(t-2L/c) dF' T(P') EF R(P') + n(t)

(1.A.14)

The above integral is performed over the target plane as

Page 25: ATMOSPHERIC PROPAGATION

qw w

Re{y(t) e-} H(f) r(t)

a) IF Model

U'A H(f)

2W 2W

- I

I U

VIF V IF

b) Normalized IF Filter Frequency Response

Figure 1.3: Heterodyne Receiver Model

Page 26: ATMOSPHERIC PROPAGATION

-26-

opposed to (l.A.10) which is performed over the receiver

plane. From (l.A.10) it can be seen that to maximize signal

return we should set F(p) = u (P)/( dpig(p)12) . In

practice, we set F,(P) = FT) to approximate this condition.

In order not to block an appreciable amount of the signal

return with H(f) we need 14 -1 1/t . To minimize the noise

passed by the filter we set W = 1/t .

In both imaging and target detection applications the initial IF

signal processing is identical. Namely, it is passed through a matched-

filter envelope-detector with output proportional to,2L/c+t~

mrij2 = l/t J 2L/ct r(t) dtf2 where r(t) is the complex envelope of2L/c

r(t) (Figure 1.3). In imaging, the scene is scanned and the complete

image built up from sequential returns Fr| 2 separated in space typically

by a diffraction limited FOV or more. In single-pulse detection ri2

is compared to a threshold where exceeding this threshold indicates

target presence. Optimal processing of returns for multi-pulse detection

also requires knowledge of Fr2 . Before the performance of these systems

can be discussed, however, propagation and targets must be characterized.

Page 27: ATMOSPHERIC PROPAGATION

-27-

CHAPTER II

ATMOSPHERIC PROPAGATION MODELS

The atmosphere, as an optical propagation medium, differs

markedly from free space. In clear weather, spatio-temporal refractive

index fluctuations are caused by random mixing of air parcels of nonuniform

temperatures. These fluctuations, called atmospheric turbulence, have a

significant effect on optical propagation. In bad weather, scattering

from aerosols, such as haze and fog, and hydrometeors, which include

mist, rain and snow, can also profoundly affect propagation.

We anticipate three atmospheric propagation effects to degrade

performance of the radar. Depending upon the relative sizes of the

phase and amplitude field coherence lengths, transmitter and receiver

apertures, and the target, different effects can become important. On

a qualitative level we have:

1. Beam Spread. In the forward path, if the transmitter

diamter exceeds the atmospheric coherence length p0, in

either turbulent or low visibility weather, random dephasing

of the transmitted field will occur. This results in a

larger target plane illuminating beam than in free space.

In Figure 2.1 the field coherence length p = (1.09 k2 C2 L)-3/50 n

under clear weather conditions is shown versus path length L

for three values of turbulence strength C2 and wavenumber kn

corresponding to 10,6 pim wavelength radiation [3]. From

Page 28: ATMOSPHERIC PROPAGATION

-28-

101

X 10.6 p.m

1 2 5 x 101 m-2/3

n

C= 5 x 10 m-n" 3

-2

2 3 4510 10 10 10

L (i)

Figure 2.1: Turbulence field coherence length p0 vs. propagation path

length L for conditions of weak turbulence

(C' = 5 x 10-16 m-2/ 3), moderate turbulence

(C2 = 10-14 m-2/3), and strong turbulence

(C2 = 5 x 10-3 m-2 3); 10.6 vim wavelength has been assumed.n

Page 29: ATMOSPHERIC PROPAGATION

-29-

this digram we see that for typical transmitted beam

diameters of 5 to 20 cm and path lengths of 1 km to 10 km

transmitter beam spread can nearly always be neglected. We

will take this to be the case. On the other hand, for

turbid atmosphere propagation, the atmospheric coherence

length p0 = (2/ s L 0' k2)2 (see Appendix A), where S'0 F

is the effective scattering coefficient and OF the forward

scattering angle, generally is much smaller than typical

beam diameters. Hence, as will be seen in more detail

later, beam spread is an important factor in bad weather,

and must be included in our analysis.

2. Scintillation. The random spatio-temporal amplitude

fluctuations due to constructive and destructive

interference of the randomly lensed light is known as

scintillation. If the target is smaller than the

amplitude coherence length of the atmosphere then the

scintillation modulates the reflected intensity. If the

target is larger than this coherence distance the

scintillation presents itself as a speckling of the

reflected radiation. In both turbulent and turbid

atmospheres this effect is significant, although, we

anticipate that signal return fluctuations will be of

different character in the clear-weather and bad weather

limits.

Page 30: ATMOSPHERIC PROPAGATION

-30-

3. Coherence loss. At the receiver, surfaces of constant

phase will become wrinkled when the receiver aperture

becomes larger than a phase coherence length (which is a

numerical factor times p 0). When this occurs optimal spatial

mode matching cannot be achieved with F*() = ET(p), resulting

in a performance degradation called coherence loss.

Referring to Figure 2.1, in turbulence, for typical receiver

diameters of 5 to 20 cm and path lengths of 1 to 10 km this

effect can also be neglected. As coherence lengths in low

visibility are normally much smaller than the receiver

aperture this effect is pronounced. More details will be

given later.

The above heuristic descriptions of propagation phenomena are

lacking in mathematical detail. We now provide these details. First,

free space propagation is described. Next, propagation through

turbulence is characterized. Finally, we discuss turbid atmosphere

propagation.

A. Free Space Model

In free space reciprocity (Eq. (l.A.8)) applies with both hLF and

hLR equal to the non-random Green's function [9]

hL(p',P) = jkL exp L - ' (2.A.1)

Page 31: ATMOSPHERIC PROPAGATION

-31-

where X is the optical wavelength, k = 27/X the wavenumber, and

superscript "o" denotes free space.

B. Turbulence Model

Without loss of generality, the stochastic Green's function for

the clear turbulent atmosphere can be represented as the product of an

absorption term, a random, complex perturbation term and the free

space result [1,31

hL(P1,P) = e- aL/2 exp[x(T',W) + j(',P)] ho(-',-) (2.B.1)

where, via reciprocity, hL = hLF = hLR. In (2.B.1), a is the

atmospheric absorption coefficient and x(I',-) ( (I',)) is the

log-amplitude (phase) perturbation of the field at transverse

coordinate p' in the z = L plane due to a point source excitation at

transverse coordinate P in the z = 0 plane. These perturbations,

before the onset of saturated scintillation, may easily be expressed

as sums of large numbers of independent random variables [1,14]. Hence,

by the central limit theorem [15], x and c are jointly Gaussian random

processes and completely characterized by their means and covariance

functions. Such a complete characterization is provided in reference

[1] but will not be given here. It will suffice to note that the mean

log amplitude perturbation obeys m = -c2 as a result of energyx x

conservation [16] where the log amplitude perturbation variance is given

by

Page 32: ATMOSPHERIC PROPAGATION

-32-

02 = min(O.124 C2 k7/ 6 L11/ 6 ,0.5) (2.B.2)x n

for Kolmogorov spectrum turbulence with a uniform turbulence strength

C2 profile.

C. Turbid Atmosphere Model

Again, without loss of generality, we can express the stochastic

Green's function as a product form

hL(,) = e -a'L/2 A(',) exp[j (',P)] h0(P',) (2.C.1)

for either the forward or return path. This representation is chosen

to emphasize that the amplitude perturbation A(p',p), and not its

logarithm, is Gaussian. This results from the fact that nearly all

of the light illuminating the target and receiver is scattered in bad

weather. Hence, these fields are the sum of a large number of

independent contributions and, by the central limit theorem [15], can

be taken to be Gaussian.

In Appendix A, the mutual coherence function (correlation

function of the atmospheric Green's function hL) is derived. This is

accomplished by considering scattering from a single particle and

assuming: (1) single scattering is concentrated about the forward

direction, (2) wide-angle scatter and back scatter can be lumped into

absorption so that the real single particle phase function can be

approximated by a Gaussian form, and (3) that L >> 1, where ' is the

Page 33: ATMOSPHERIC PROPAGATION

-33-

modified atmospheric scattering coefficient, so that the direct

(unscattered) beam contribution to the field is insignificant compared

to the scattered field.

The result we use is [2,17,18]

<hL (Til9 ) h*(p )> = e a exp[-(j 2 + IT' - p 2 +

(T1 _-P _2))/3pf] ho(TP ,p) h*( , 2 (2.C.2)

In (2.C.2), '(> Ba) is the modified absorption coefficient containingaa

wide angle scatter and back scatter contributions and

Po= [L e2 k2 (2.C.3)

is the atmospheric coherence distance where e F is the root-mean-square

forward-scattering angle of the Gaussian single particle phase function.

Furthermore, the variance of the phase perturbation is large enough

to say that hL(P',P) is zero mean and <hL(P',Pl) hL(' P2 )> =

The above model is known as the multiple forward scattering

(MFS) approximation. The validity of the model is not fully

established. It should yield correct results when each scattering

particle is significantly larger than a wavelength X. In this case,

the true single particle phase function would be highly peaked about

the forward direction so that the previous assumptions would apply.

Page 34: ATMOSPHERIC PROPAGATION

-34-

As discussed in Appendix A the correlation function (2.C.2)

accounts only for scattered light. That is, results derived from

(2.C.2) disregard the unscattered portion of the beam. In order to

take this into account the unscattered beam is taken to be the

free-space result reduced by the extinction (absorption and scattering)

loss. In this thesis, the main theoretical development is aimed at

making use of the scattered light in the context of an optical radar.

Clearly, before use can be made of the scattered power, as calculated

via the MFS theory, it must dominate the extinguished free-space power.

More will be said about this issue in Chapter VII.

D. Backscatter

When the monostatic radar, as described in Chapter I, is used

in inclement weather there may be significant backscatter return from

the hydrometeors and aerosols present in the propagation path. Clearly

this return is undesirable in imaging and target detection applications,

but unavoidable in such weather conditions. To see when such a

return can be significant its power must be compared to those of the MFS

and extinguished free-space target returns. The backscatter

contribution to the radar return will be examined in Chapter IV,

Section H and again in Chapter VII.

Page 35: ATMOSPHERIC PROPAGATION

-35-

CHAPTER III

TARGET INTERACTION MODEL

Here we discuss how the reflected radiation is related to the

target illumination. We first discuss the planar target model used in

this analysis. This model is then related to the more usual bidirectional

reflectance in the following section.

A. Planar Reflection Model

In scalar paraxial optics reflection of an optical beam from a

spherical mirror is generally represented in terms of a planar

reflection model. If the incident and reflected fields travelling

nominally along the z-axis are noted ut (',t) and ur(P't), respectively,

we have the relation

u- (I',t) = t(F',t) r expr-jkjp' - 1cf2/Rcl (3.A.l)

where r is the intensity reflection coefficient, Rc is the radius of

curvature of the mirror and p c is the transverse location of the center

of curvature. The use of (3.A.1) presupposes -c lies on or near the

z-axis and Rc is much larger than the beamwidth of u-t More generally

we might represent a polished reflecting surface by incorporating into

(3.A.1) spatially varying intensity reflection coefficient, radius

of curvature and center of curvature. That is

Page 36: ATMOSPHERIC PROPAGATION

-36-

ur(W',t) = ut(',t) r- (') exp(-jk I' - (P)I 2 /Rc(p')) (3.A.2)

would be our model. In accordance with (3.A.2) we shall assume that

for all targets of interest a planar target interaction model (l.A.6)

is applicable

Ur(I't) = u-t(p',t) T(P') (3.A.3)

In general, T(P') will contain two components, the so-called

specular (glint) and diffuse (speckle) reflection components. We

express this as

T(P') = eJO T (p') + Ts(p') (3.A.4)

The glint component, T (P'), is nonrandom and represents the component

of the reflected liqht that is due to the smoothly-varying target shape.

This component may be described by (3.A.2) or (3.A.1). The random

phase e is assumed to be uniformly distributed over [O,2tr] and represents

our uncertainty of target depth on spatial scales on the order of a

wavelength. On the other hand, the speckle component T (_') is random

and represents that part of the reflected light that is due to the

microscopic surface-height fluctuations of the target. This component

may be assumed to be a Gaussian random process with moments [19-21]

Page 37: ATMOSPHERIC PROPAGATION

-37-

<T ( ')> = 0 (3.A.5)

<T (() ) Ts(T) )> = 0 (3.A.6)

-s 1 s 21 )6~<TS a' _S )> = X Ts( ) 6( - ) (3.A.7)

Use of the above moments is justified by the fact that a purely

diffuse target would turn a perfectly coherent illuminating beam into

a spatially incoherent reflected beam. The quantity Ts (p') can be

interpreted to be the mean-square reflection coefficient at p'.

What is sought in operation of the radar is information about

the target. In terms of the preceding model it is worthwhile to

mention what information we are seeking. We regard T(P') as the target.

For the glint component we are interested in the field reflection

coefficient (P') while for the speckle component we want information

on the mean-square reflection coefficient T(P'). Note that neither

the random phase e nor the exact speckle component field reflection

coefficient Ts(P') is regarded as interesting.

B. Relationship to Bidirectional Reflectance

Let us examine the relationship between the preceeding target

statistical model and bidirectional reflectance, the target signature

quantity that is generally measured [3,8,22]. For the target geometry

Page 38: ATMOSPHERIC PROPAGATION

-38-

of Figure 3.1, the bidirectional reflectance may be defined as

(X; ; = 1 < dp' exp1[2f - ) -P '] T(') 12> (3.B.1)1 r X2AT - r

where AT is the target's projected area. If the input field is the

plane wave exp(j27 i -p') then the total reflected field is given by

exp(j2r 7 -p') T(p'). The portion of this reflected field that is in

the fr direction is the function-space projection [23] of

exp(j2T i -P') T(I') onto the function exp(j27 Ir -p'), that is, the

Fourier transform of the reflected field. Hence, the bidirectional

reflectance gives the ratio of the average reflected radiance (W/m2sr)

in the direction fR to the incident irradiance (W/m2) propagating in

the direction .

Page 39: ATMOSPHERIC PROPAGATION

W W W ww

REFLECTED FIELD U r(U,+)r _ TA RG ETKiiiiiiii fr

TARGET-PLANE FIELD ' z

INEFFECTIVE PLANE OF INTERACTION

z L

Figure 3.1: Geometry for defining bidirectional reflectance p'(x;Ti,Tr); the target plane field ischosen to be a plane wave of wavelength x propagating in the direction of the unit vectorii (Xfi is the projection of ij on the z = L plane); the radiance of the reflected fieldis measured in the direction of the unit vector ir (xfr is the projection of ir on thez = L plane).

4W MW 1W W

Page 40: ATMOSPHERIC PROPAGATION

-40-

CHAPTER IV

SCANNING-IMAGING RADAR ANALYSIS

With atmospheric propagation and target interaction models in

hand we are ready to proceed with the scanning-imaging radar analysis.

This will consist mainly of a signal-to-noise ratio (SNR) analysis.

To a lesser extent resolution and correlation of simultaneous target

returns are also considered.

We assume that an image is built up through scanning a scene

diffraction-limited FOV by diffraction-limited FOV. With this kind of

imaging system successive returns from the same direction are

separated in time typically by tens of milliseconds. Since atmospheric

correlation times in both turbulence and low visibility are considerably

shorter than 10 msec, successive returns can be taken to be independent.

Accordingly, with the SNR definition given below, the N-pulse,

single-pixel SNR is N times the single-pulse SNR. Coupling this with

the fact that speckle target returns from disjoint diffraction limited

FOV's are independent when the MFS model applies, as shown below, the

single-pulse SNR is a reasonable performance measure.

We begin this chapter with a formulation of the single-pulse

SNR problem. In the following section, speckle target resolution in

bad weather is considered. We then discuss how the various atmospheric

degradation effects will be identified in the SNR formulas. Following

Page 41: ATMOSPHERIC PROPAGATION

-41-

this, SNR results in turbulence and low visibility will be presented.

We then explore the SNR degradation effects of beam wander due to

reset error in the radar aiming mechanism or atmospheric beam steering.

Next, correlation of simultaneous returns from different directions is

considered. Finally, backscatter from particles in the propagation

path is examined.

A. Sinale Pulse SNR

We are interested in the target reflection strength and

accordingly consider

2L/c + tp

Ir12 = 1l/t J r(t) dt| 2 (4.A.1)

- 2L/c

the output of a matched filter envelope detector with input r(t).

Ignoring the passband filter of Figure 1.3 and assuming tp = 1/W (the

tp second integration has approximately the same effect on the IF signal

r(t) as the filter) we have

r12 = Ix + n12 (4.A.2)

where the signal return is given by the target plane integral

j'ft pP T'~ (4A3x = d' T(p') G(p') s(i')i lyAen

and the noise n is zero-mean, complex-Gaussian, statistically i-ndependent

Page 42: ATMOSPHERIC PROPAGATION

-42-

of x, with moments

<n2> = 0

<In 2> = 1

(4.A.4)

(4.A.5)

Note that the normalization chosen in (4.A.3) leads to the simple

result (4.A.5). The mean of the observation (4.A.2) is

_= 2> + <In12> (4.A.6)

The term <In! 2> is signal independent and due to receiver noise. Hence,

we define the image signal-to-noise ratio to be

(<lr2> - <ln2>)2SNR = _

Var(I r12)

<Ix 12>2

Var( I r12)(4.A. 7)

That is, the ratio of the square of the signal portion of the observation

mean to the observation variance. Since n is zero mean complex-Gaussian

with statistics (4.A.4), (4.A.5) it is straightforward to show

SNR = CNR/2CNR/2 11 SNR SAT 2CNR

(4.A.8)

where the carrier-to-noise ratio (CNR) has been defined to be the ratio

of the signal portion of the observation mean to the noise portion

Page 43: ATMOSPHERIC PROPAGATION

-43-

CNR (4. A.9)<tn_ 2>

and the saturation SNR is

<|x|2>2SNRSAT = <I.X12> (4.A.10)

AT Var( xi2)

For CNR 5 or CNR 2 (10 SNRSAT we can disregard the last term in

the denominator of (4.A.8) and obtain

SNR = SNRSAT + CNR/2 SNRSAT (4.A.ll)

The maximum value of SNR is, from (4.A.ll), seen to be SNRSAT and is

achieved when CNR >> 2 SNRSAT. Physically, this limiting SNR results

when noise fluctuations become negligible in comparison to signal

fluctuations. From (4.A.8) and (4.A.ll) it can be seen that to complete

the analysis we need to evaluate SNRSAT and CNR for a number of

atmospheric/target situations. In order to know these quantities we

must calculate <tx!2 > and Var(tx1 2 ). Direct calculation of <tx 2 > via

the definition (4.A.3) is not too difficult but use of the same approach

for Var(tx1 2 ), while in principle straightforward, yields results which

quickly become unmanageable. A more reasonable approach is to try to

develop a complete statistical characterization for Ix12 for a number of

atmospheric/target scenarios. This approach, besides being simpler,

has the advantage of providing the information needed for the target

Page 44: ATMOSPHERIC PROPAGATION

-44-

detection problem. This is the approach to be taken.

Up to now no assumption has been made regarding the statistics

of the signal return x. Therefore, the formulation of the imaging SNR

problem in equations (4.A.1) to (4.A.ll) is equally applicable to free

space, turbulence, and low visibility as well as beam wander induced

fluctuations. To finish this section we give selected free space

SNRSAT and CNR results.

Here we shall assume a Gaussian-beam system described by

FT(P) = F*(P) exp(_IP12/2p2) (4.A.12)(TrpTY

where PT is the radar transmitter and receiver pupil radius.

For a resolved speckle target (i.e. one that is larger than the

illuminating beam size) with intensity reflection coefficient r in free

space it is simple to show that

T.1P t 2TrP2CNR = 0 r (4.A.13)

ho L

and [3]

SNRSAT = 1 (4.A.14)

For an unresolved glint target, with field reflection

coefficient

Page 45: ATMOSPHERIC PROPAGATION

-45-

T (p ) = exp[- L' 1 2/2r2] (4. A.15)

-<s

we have

CNR = L4rrsPTj (4.A.16)

and a saturation SNR equal to infinity. More detailed examples of

this type will be given later.

B. Speckle Target Resolution in the Low Visibility Atmosphere

We consider the output of the matched filter envelope

detector (4.A.2) when the noise n is negligible so that I 2 T2

where

x(fT hjtp PT d-p' T(p) Eg p') Flp) (4.B.1)

and the, atmosphere is characterized by (2.C.2). In (4.B.1) we have

explicitly noted that the signal return is a function of pulse

propagation direction as defined by TT (-see below). Limiting

ourselves to pure speckle targets, the mean signal return < x(-T 2>

will be considered. In examining this quantity we will be concerned

with what portion of the target contributes to the signal return.

Page 46: ATMOSPHERIC PROPAGATION

-46-

Before giving the expression for the mean signal return it is

worthwhile to develop some preliminary results. Namely, the field

spatial correlation functions <F(Cj) gp)> and qR

should be found. For the resolution issues discussed in this section

it is necessary to know these correlation functions only for p= p.

The more general result ( 5 is useful in the sequel, though,

and will be given here. To facilitate the calculations Gaussian

transmitted and local-oscillator spatial modes are assumed

FT(P) exp[- I 2/2 p2 + j 2 T f-] (4.B.2)T (~TrP 2 )T

(T

F*(P) -exp[- P 1 2 /2 p + j2nTr - ] (4.B.3)ex E-rP/ 2 )1f2 R(RR

In the above equations PT and PR correspond, respectively, to the

transmitter pupil radius and receiver pupil radius and fT determines

the direction of propagation as this pulse will illuminate a circular

region in the z = L plane centered on the transverse coordinates

P = ALfT.

The atmospheric Green's functions hLF(P',P) and hLR(PP') are

taken to be statistically independent. This, as stated earlier, is

the "effectively bistatic" assumption. In order for this assumption to

be reasonable when the radar is in a monostatic configuration (i.e. when

the transmitter and receiver are colocated) the roundtrip delay for

Page 47: ATMOSPHERIC PROPAGATION

-47-

pulse propagation to the target and back to the radar must be longer

than the atmospheric coherence time Tc. An upperbound on this coherence

time [24,25] can be taken to be the time for a frozen atmosphere moving

at the transverse wind velocity Vt to move a coherence distance p ,

i.e. Tc < P0/Vt. This expression ignores random motion of the air

molecules which could decorrelate the atmosphere much more quickly than

the preceding expression would suggest. For T' = s'L = 10,

forward scattering angle eF = 10 mrad, 10.6 ym radiation and transverse

wind velocity V t = 10 km/hr this upper bound on the atmospheric

coherence time is p0/V t 25 psec. The roundtrip delay time td = 2L/c

is 6.7 -psec at range L = 1 km so that delay is equal to the upper

bound when range is approximately 3 km. Even with the maximum value

for coherence time the "effectively bistatic" assumption can be made

for reasonable target- ranges.

The spatial modes (4.B.2), (4.B.3) and the definitions (l.A.12),

(l.A.13) in combination with the moment (2.C.2) are used to develop

expressions for the field spatial correlation functions <L 4$>, <L 4>.

To facilitate interpreting the results of later sections, the

atmospheric coherence distance p0 is taken to be different for the

forward and return paths. Specifically, poF replaces p0 in (2.C.2) for

hLF(P',p) while poR replaces p0 for hLR(p',). Physically, such an

assumption is unreasonable for a monostatic radar and is made only to

aid interpretation. The correlation functions are then given by

Page 48: ATMOSPHERIC PROPAGATION

-48-

-'Le 2 exp[-p'TrrbF

ALI 2/rbF

- exp[-p |2 /rcF2 exp[j k (p' - XLfTT) ]exp[j2pT'SoF c T d

(4. B. 4)

where for convenience we have expressed the result in sum and difference

coordinates

(4. B.5)

(4.B.6)

In Equation (4.B.4) the beam radius rbF is

rbF = (AL/pF)2 + (AL/2TrpT) 2 +

the coherence distance r cF is

(4.B.7)

2 2

bF "oF

TT (AL/hrpOF)2 + +(AL/2pT) + pT +

rc = (4.B.8)

<_-F( ') -LPF ) 2

-c = 1 (p - + p )

+ p2F /4

Page 49: ATMOSPHERIC PROPAGATION

-49-

and the phase radius of curvature RoF is

4p + 3pF 2 rbFR = L[T + bF (4.B.9)oF P + 3p F r 2 T

The corresponding expressions for <-RC) L*(p)> are found from

(4.B.4) - (4.B.9) by replacing pT by PR and PoF by PoR everywhere so

that r F becomes rbR, rcF becomes r2R, and ROF becomes ROR.

Assuming the radar is "effectively bistatic," that it uses

the Gaussian spatial modes (4.B.2), (4.B.3) with PT = PR and poR= oF Po

and that the target is pure speckle with moments (3.A.5) - (3.A.7),

the mean signal return is

< X( T)h T e-2 L x-- d~p' T (-P') exp[-Ip' - XLfT 2 /r es<' 0 7T 2>rh4 s ee5

(4.B.10)

where rbF = rbR r and

1 ~

re r /2 = [(XL/2)rpT 2 + pT] 1 + 2 (4.B.11)(AL/2rfTP + PT

is the target-plane resolution e~1 spot size. Equation (4.B.10) can be

interpreted as saying the signal return is an average over a spot of

Page 50: ATMOSPHERIC PROPAGATION

-50-

area r es centered at Lf If the target is in the radar's far field

this becomes

rres 2 (XL/2mpT) 2 [1 + ( T (4.B.12)

which can be interpreted as saying that the signal return is an average

taken over approximately 1 + 4/3 (p T/P)2 diffraction-limited fields of

view.

C. Identification of Atmospheric Effects

Although we discussed earlier a number of atmospheric effects

expected to degrade the performance of an optical radar, no indication

was given as to how these effects might be identified in our performance

analysis. Here we discuss how this identification will be made in the

SNR and CNR formulas to follow.

1. Forward-path beam spread loss. If PT is the transmitted

beam radius and p0 the atmospheric coherence distance, any

degradation due to beam spread in the forward path can be

eliminated by letting PT become small in comparison to p0 .

Under this condition diffraction would dominate any

atmospherically induced beam spread and this loss mechanism

should be eliminated from the CNR and SNR formulas. Note

that this approach will be useful only in interpreting

low-visibility results as beam spread was already shown to

be unimportant in turbulence.

Page 51: ATMOSPHERIC PROPAGATION

-51-

2. Return path beam spread loss. After reflection of the

illuminating beam the phase fronts of the return beam will

undergo additional wrinkling and hence additional beam spread

will be incurred. This source of beam spread loss is not

as easily identifed in the equations as the forward-path

loss. But it should not be present when the target is pure

speckle as the reflected light is effectively radiating

into the 27r steradians solid angle in front of the target,

nor should it be present when the target is pure glint and

smaller than the coherence distance p as diffraction then

dominates beamspread.

3. Receiver coherence loss. As this loss mechanism is due to

a wrinkled phase front in the received field it can be

eliminated by decreasing the receiver pupil radius PR to

satisfy PR < Po. In this case, the received phase front

will be flat over the receiver aperture and the spatial

modes of the received and local oscillator fields will

match. Coherence loss is not important in turbulence, as

discussed earlier, so this approach is useful only in low

visibility.

4. Scintillation. The effects of scintillation will be

difficult to see in the CNR and SNRSAT formulas. These

effects will become apparent when complete statistical

characterizations of target returns are presented.,

Page 52: ATMOSPHERIC PROPAGATION

-52-

We now proceed with a series of SNR, CNR examples corresponding to

different atmospheric/target scenarios. Turbulence results will be

presented first, followed by low visibility results.

D. Turbulence SNR Results

The results cited in this section have previously appeared

in references [3-5,26]. They assume the Gaussian-beam system

(4.B.2), (4.B.3) with perfect transmitter, local oscillator mode

matching Fp) = F*(p), PT = PR. We present these results as a series

of examples.

Case 1. Turbulence, Unresolved Glint Target

Here a pure glint target T(p') = T (p') e j that behaves like

a spherical reflector over the illuminated region

T (p') = F exp(-jkjp'-'|/Rc) p' -ALfT (4.D.l)c_ c T d

and satisfies

Rc << L (4.D.2)

(XRc 0 (4.D.3)

is assumed. Equation (4.D.2) should often if not always hold. It can

Page 53: ATMOSPHERIC PROPAGATION

-53-

be shown that the effective radiating region of the target (4.D.1)

(i.e. the region within the illuminated portion of the target that

makes an appreciable contribution to the target return) has nominal

diameter (R c)P. Hence (4.D.3) amounts to saying that the effective

radiating region is smaller than an atmospheric coherence area. A

target satisfying (4.D.l)-(4.D.3) is called a "single glint" target.

Furthermore, assuming that the target lies entirely within the

illuminating beam, i.e. it is unresolved, we find that the signal

return can be expressed as

- 4cr 2

1x_12 = CNRgu e X exp[4X(p ,0)] (4.D.4)

where p' is the glint reflection point of (4.D.1) and exp[4X(',U)]

is a lognormal random variable. In (4.D.4) CNRgu is the unresolved

(denoted "u") glint (denoted "g") target, tubulent atmosphere

carrier-to-noise ratio and a2 is the variance of the log-amplitudeX

perturbatiort X( 9,9) .

It follows from (4.D.4) that the single pulse image SNR

satisfies

CNR /2SNR u gu (4.D.5)

g 1 + CNR gu(e X - 1)/2 + 1/2 CNRgu

For CNRgu z 5, Equation (4.D.5) takes the standard form (4.A.ll) with

Page 54: ATMOSPHERIC PROPAGATION

-54-

SNRSAT 1 (4.D.6)SAT 16 a2gu e X

If no turbulence is present, we have a2 0 and the saturation SNRX

(4.D.6) is infinite as predicted by (4.A.10). For a2 > 0, SNRX gu

initially increases with increasing CNR until it reaches the

scintillation-limited value (4.D.6). For cr2 > 1/16, SNRSAT isX -STgu

severly limited by turbulence. Multiframe averaging is required to

overcome this limit.

Case 2. Turbulence, Speckle Target

When the target is assumed to be pure speckle, T(p') = T '

the single pulse image SNR satisfies

CNR /2

SNRs 16 52 (4.D.7)

1 + CNR s[1 + 2(e X -1)C]/2 + 1/2 CNRs

where C is the log-amplitude aperture averaging factor [16,41] given by

the approximate expression

4 P2 /XL4T~ 2 (4.D.8)1 + 4 p T //L

in the weak perturbation regime and CNRs is the speckle (denoted "s")

target, turbulent atmosphere CNR. If the mean-square reflection

coefficient T(p') does not vary appreciably over the region illuminatedS

Page 55: ATMOSPHERIC PROPAGATION

-55-

by the radar (i.e., the target is resolved) the target return can be

expressed as

Ix 2 = CNRsr v e2u (4.D.9)

where v is a unit mean exponential random variable and u is a Gaussian

random variable, statistically independent of v, with mean -a2 and

variance a 2 satisfying

2 1e4c - 1 = C(e

In (4.D.9), v represents the target speck

The probability density function for the

w = v e2u is given in [27]. If CNRs z 5,

standard form (4.A.ll) with

SNRSAT =5As

6a2X - 1)

(4.D.10)

le and u the scintillation.

unit mean fluctuation

Equation (4.D.7) takes the

1

16 a21 + 2(e X -_1)c

(4.D.11)

From (4.D.11) note that SNRSAT < 1 with SNRSAT = 1

speckle limited saturation SNR.

For more details on turbulence SNR results,

examples, the reader should see [3-5,26].

corresponding to

including system

Page 56: ATMOSPHERIC PROPAGATION

-56-

E. Low Visibility SNR Results

In this section first CNR and then SNR results will be

developed for bad weather radar operation. Again, the Gaussian

transmitted and local oscillator spatial modes (4.B.2), (4.B.3) as

well as an atmosphere characterized by (2.C.2) are assumed. The

correlation function (4.B.4) then applies to this case where we will

take the coherence distances for the forward and return paths to be

different poR PoF* This will aid in interpreting the results of this

section.

Assuming that the atmospheric coherence time Tc is short

enough to justify saying the radar is "effectively bistatic" the CNR

(4.A.9) becomes

-nP tCNR = Pt dpi dp <T(T ) T*(T)><-F ) L (Pj4

0 J (4.E.l)

Our task is now reduced to evaluating the integral (4.E.1). This is

done as a series of examples, each corresponding to a different target.

We will interpret the results in terms of the previously mentioned

atmospheric degradation mechanisms.

Case 1. Low Visibility, Unresolved Speckle Target

The target is pure speckle T(-') = T (p') with-P -S

Ts(P') exp[-I' 2/r' ] (4. E. 2)

Page 57: ATMOSPHERIC PROPAGATION

-57-

The mean square reflection coefficient Ts is Gaussian with width rs

and centered on the z axis. Although no real target would have the

form (4.E.2) it is chosen to allow closed form evaluation of (4.E.1).

Also the CNR results below would be applicable to any unresolved

speckle target if we replace 7rr with the speckle target area AT.

Assuming

rs<< rbF9rbR (4.E.3)

i.e., that the target is unresolved by the radar, the carrier-to-noise

ratio (4.E.1) becomes

CNR = CNR 0 exp[-2'L]su su 1/3(XL/pOFTO 2 1/3(XL/p oR7 2 a

1 + 1 +

(XL/2T) 2 + p (XL/2rpR + R

*exp LTI 2 (4.E.4)

rbF rbR/ rbF + rbR

where CNR 0 is the free space unresolved speckle target CNRsu

TIP t x2 r 2

CNR 0 = T _ s r5 (4.E.5)su hv 0 ' [(XL/21pT) 2 + p2][(XL/2pR)2 + pR]

Page 58: ATMOSPHERIC PROPAGATION

-58-

The second multiplicative term in (4.E.4) is the forward-path beam

spread loss since it approaches 1 as PT +* 0. Similarly the third term

can be identifed as the coherence-loss term as it approaches 1 as

PR - 0. The fourth term represents absorption and the last gives the

loss due to the misalignment of the radar, i.e., that the center of

the illuminating field is IXLfT1 from the target center. Note that no

return-path beam-spread loss is evidenced by (4.E.4), as predicated

earlier. It should also be noted that the total free space beamwidth

(XL/2pT) 2 +p is used instead of the far field approximation

(XL/27rrpT) 2. At C02 wavelength 10.6 ypm and transmitter pupil radius

PT = 6.5 cm, the change over from near field to far field occurs at

approximately 2.6 km, midway through the expected useful range of the

radar.

Case 2. Low Visibility, Resolved Speckle Target

The target is again pure speckle T(P') = T (-') with mean

square reflection coefficient (4.E.2). Assuming

rs >> rbF, rbR (4.E.6)

so that the target is resolved, the CNR becomes

Page 59: ATMOSPHERIC PROPAGATION

-59-

0CNR sr= CNR sr

+ 1/3(XL/poRr) 2 1/3(AL/pOF7F)2

(XL/2R)2+p2+( L/2r)2+P (XL/2rpR)2+p2+( L/2 pT) 2+p2

exp[-2S;L] exp[-fXLfT 2 /r2 ] (4.E.7)

where CNR0 is the free space, resolved, speckle target carrier-to-noisesr

ratio

CNRP0 - Tt p 1sr hv0 'T (XL/2TpT) 2 + p + (AL/2pR)2 + PR

(4.E.8)

The final two terms in (4.E.7) are identifed as absorption and

misalignment losses as before. The forward-path beam spread and

coherence loss are given by the second term. The second term in the

denominator of the quantity

1+ 1/3(XL/poRr) 2

(XL/2TrpR) 2 + p + (AL/2rp) 2 + p2

+ 1/3(XL/poF')

(AL/27rpR) 2 + p2+ (AL/27rpT) 2 + 2

(4.E.9)

gives the coherence loss while the last term gives the forward-path

beam spread loss. We see that forward-path beam spread loss in the

Page 60: ATMOSPHERIC PROPAGATION

-60-

resolved case is affected by both PT and PR but affected only by PT

in the unresolved case equation (4.E.4). In order to understand the

difference in these two cases, consider the situation when the return

medium is free space p = . Clearly, there can then be no coherence

loss. From Figure 4.la it can be seen that in the unresolved case only

the on-target (essentially constant) incident power density is affected

by beam spread. The spatial mode of the reflected light depends on

target shape and not on forward path beam spread. Hence, changing PR

cannot reduce forward-path beam spread loss as the spatial mode of

the reflected light is independent of the beam spread. In Figure 4.lb

(resolved case) we have the situation in which the reflected light has

a spatial mode that depends upon forward-path beam spread and all of

the incident power is reflected. By choosing PR large enough, any

loss associated with this beam spread can be eliminated. That coherence

loss is affected by PT in (4.E.7) can be similarly explained.

Case 3. Low Visibility, Unresolved Flat Glint Target

The target is pure glint T(p') = ejeT (p') where

T (j')= p2 exp[-IK'12 /2r2] exp _ 2 (4.E.10)

The intensity reflection coefficient T(p)1 2 is Gaussian with width rs

as previously. We assume that the target is flat so that the phase

radius of curvature

Page 61: ATMOSPHERIC PROPAGATION

-61-

Beamspread

Target

RADAR 2rbF

Incident powerconstant over

----...target

Free Space Beamwidth

Reflected spatial mode is independentof forward-path beamspread

a) Unresolved Target

.... Target

Beamspread

RADAR 2r F

Free SpaceBeamwdith

Reflected Spati'_'FMode depends onforward-path beamspread

b) Resolved Target

Figure 4.1: Reflected Spatial Modes from Resolved and Unresolved SpeckleTargets.

Page 62: ATMOSPHERIC PROPAGATION

-62-

Rg

= O

We further assume the target is unresolved

rs << rbF, rbR

the radar is in the target's far field

r << a

and the laser is aimed directly at the

fT = 0

(4.E.13)

target

(4.E.14)

so that the CNR becomes

CNR = CNR 0gu gu1/3(AL/rpoF)2

(AL/27rpT)2 + p

1/3(L/TrpoR )2

(XL/2rpR) 2 + 12R

-1

4r 2 4r2

1 + +r 2 r 2rCF rcR

exp[-2 'L] (4.E.15)

is the free space unresolved flat glint target CNRwhere CNRgu

(4.E.11)

(4.E.12)

Page 63: ATMOSPHERIC PROPAGATION

-63-

nP t 4r'CNR0 - tp 4s (4.E.16)

gu hv0 [(XL/2rrpT) 2 + p{1[(AL/2TrpR)2 + PI]

Since the effective radiating region of a spherical glint target has

nominal diameter (AR ) the above results can be applied to the case

of finite R by replacing 2rs with (AR )2. In (4.E.15), the second,

third and fifth terms are recognized as forward-path beam spread loss,

receiver coherence loss and absorption loss, respectively, as in

(4.E.4). The fourth term can be interpreted as return-path beam spread

loss. Justification of this last identification follows from the

fact that this term approaches 1 as rs -+ 0. To see more clearly what

is transpiring, note that this term is given by

(4.E.17)r2 r2

1+ +2 2

PoF PoR

when the target is in the transmitter's far field and PR > PoR'

PT > oF. Since a coherent radiator in the target plane of size rs

would have beam size 1/3(XL/p oR)2 + (XL/27r s)2 in the far field, it

is clear that beam spread would affect beam size only when rs >oR'

which is exactly when the third term in the denominator of (4.E.17)

becomes significant. To see how low visibility in the forward path

would affect return-path beam spread consider Figure 4.2. In Figure

4.2a an essentially flat phase front is reflected from the target

Page 64: ATMOSPHERIC PROPAGATION

-64-

Incident Phasefront

Reflected Phasefront

ii2r

Target

a) Target smaller than a coherence area, rs < rcF : PoF

Incident Phasefront

Reflected

2rs

Phasefront

arget

b) Target larger than a coherence area, rs > rcF ~ oF

Figure 4.2: Reflected Phasefronts from Glint Targets

.

Page 65: ATMOSPHERIC PROPAGATION

-65-

while in Figure 4.2b a wrinkled phase front is reflected. The

wrinkled phase front would clearly cause beam spreading after

reflection. This beam spreading becomes significant when rs >oF;

this is exactly when the second term in (4.E.17) becomes significant.

We now turn our attention to developing complete statistical

characterizations of the single-pulse target return. These

characterizations will be of the form (4.D.4), (4.D.9) wherein the

signal return |x 2 is represented as a product of constants and

independent random variables. This is most easily accomplished via a

series of examples.

Case 1. Low Visibility, Small Speckle Target

Note that the target has been designated as "small" instead

of "unresolved" as in the CNR examples. The reason for this is the

requirement on the size of the target is different here than previously,

as will become apparent. Describing the target T( ') = T (P) by

(4.E.2) we express the signal return (4.A.3) as

x_= p)T dp' Ts P 2P F (4.E.18)

where we have made the definition

dP' ! (P') L.( ' EF( 'tA J(4.E.19)

(f dp' Ts(-') 1 12 F

Page 66: ATMOSPHERIC PROPAGATION

-66-

Assuming T(p') is a Gaussian random process it is a simple matter to

show, by consideration of the conditional moments of c given (R and F

that c is a zero mean complex Gaussian random variable with independent

identically distributed real and imaginary parts and variance

< 2> = 2 (4.E.20)

Assuming that the target is small compared to a coherence area

(4.E.21)r2 << rF2 p 2

< cF PoF

s cR OoR (4.E.22)

we can then say

(4.E.23)

__F I (. 2 (4.E.24)

inside the integral (4.E.18). It then follows that the signal return

can be expressed as

x 2 = CNR u v w (4. E. 25)

Page 67: ATMOSPHERIC PROPAGATION

-67-

where u, v, and w are independent, identically distributed, unit mean,

exponential random variables. In (4.E.25), u represents target

fluctuations and v, w the forward and return path atmospheric

fluctuations. These atmospheric fluctuations represent the randomness

imposed by the terms (4.E.23), (4.E.24).

Two points need to be mentioned relating to the assumptions

(4.E.21) and (4.E.22). The first is that any target satisfying these

conditions is necessarily unresolved as rcF < rbF and rcR < rbR'

Second, is that this condition is an extremely stringent one as poF and

poR may easily be less than a millimeter.

The saturation SNR for this case is now easily calculated

from (4.E.25)

SNRSAT <X 2>2 (4.E.26)ss Var( xf2 ) 7

where "ss" denotes small speckle target. A physically more realistic

situation is now presented.

Case 2. Low Visibility, Large Speckle Target

Again, the target T(p') = T (') is described by (4.E.2) so

that (4.E.18) applies where E is a complex Gaussian random variable.

Assuming that the target is large compared to a coherence area

Page 68: ATMOSPHERIC PROPAGATION

-68-

r2 >> r2 (4.E.27)s cF

r2 >> r2 (4.E.28)s cR

we can then argue there is sufficient target plane aperture averaging

so that the integral (4.E.18) becomes nonrandom and can be replaced by

its mean. Taking this to be the case we then have

jx_| = CNR u (4.E.29)

where u is a unit mean exponential random variable representing target

fluctuations.

The conditions (4.E.27), (4.E.28) do not impose a very

stringent limitation on the target size. In fact, the target can be

either resolved or unresolved and still satisfy them. The saturation

SNR is now given by

SNR SAT = 1 (4.E.30)

where "sZ" denotes large speckle target. The cases of small and

resolved glint targets are now presented.

Page 69: ATMOSPHERIC PROPAGATION

-69-

Case 3. Low Visibility, Small Glint Target

The target T(p') = e a T (P') is given by (4.E.10) so that

the signal return is

[nt P T

x_= h TeJe d' T (T') E-RP) -F(E.)3

If we assume the target is small, i.e.

s a cF

r2 << rcR

(4.E.32)

(4.E.33)

then we can replace g(P') and ') in the above integral by (

and Lg(U), respectively, as in the small speckle target case.

follows that

fx 2 = CNR v w

It then

(4.E.34)

where v and w are independent identically distributed unit mean

exponential random variables representing forward and return path

fluctuations, respectively. The saturation SNR for this case is

1SNR SAT g (4.E.35)

(4. E.31 )

Page 70: ATMOSPHERIC PROPAGATION

-70-

The same comments regarding the implications of (4.E.32), (4.E.33)

apply here as in case 1.

Case 4. Low Visibility, Resolved Glint Target

For a glint target T(P') = ej T (P') specified by (4.E.10)

that satisfies

r 2>> rb 2r > bF

s bR

(4.E.36)

(4. E.37)

it can be easily argued, by considering an unfolded geometry for the

radar configuration Figure 1.1, that the received field u (5,t) is

Gaussian. It then follows from Eq. (l.A.10) that x is Gaussian and

|x2 exponential

1x1 2 = CNR v (4.E.38)

so that

SNRSAT gr= 1 (4.E.39)

This complete the single pulse signal-to-noise ratio analysis

for the "effectively bistatic" radar. In the next section, the effects

Page 71: ATMOSPHERIC PROPAGATION

-71-

of beam wander on unresolved target returns is considered.

F. Beam Wander Effects

The power reflected from an unresolved target will fluctuate

due to beam wander effects from radar pulse to radar pulse. The

source of the beam wander may be actual reset error in the aiming

mechanism of the radar or beam steering effects of turbulent atmosphere

propagation [1]. To account for these fluctuations in our model the

expressions for the matched-filter envelope-detector output IX| 2

(i.e. (4.D.4), etc.) are multiplied by another unit mean random

variable. If we let the unit mean random variable w represent the

collective effects of target/atmospheric fluctuations (excluding beam

steering) and the unit mean random variable v represent beam wander

induced fluctuations then we can say in general

IX 2 = CNR w v (4.F.1)

Taking w and v to be independent the saturation SNR is

SNRSAT = (4. F.2)SAT =Var(w) Var(v) + Var(w) + Var(v)

The expression reduces to

SNRSAT = SNRSATbw (4.F.3)

Page 72: ATMOSPHERIC PROPAGATION

-72-

when

SNRSATta >> SNRSATbw

SNRSATta

SNRSATta

(4.F.4)

(4.F.5)> 1

Var(w) (4.F.6)

is the saturation SNR when only target/atmospheric fluctuations are

present and

(4.F.7)SNRSATbw Var(v)

is the saturation SNR when only beam wander fluctuations are present.

By extension, when the conditions (4.F.4), (4.F.5) are satisfied we

can say

JxJ 2 = CNR v

and

where

(14.F. 8)

Page 73: ATMOSPHERIC PROPAGATION

-73-

or equivalently

IxI = <lxI> z

The random variable z is the unit mean fluctuation on

related to v by

lxi and is

V (4.F.10)

We develop statistical models for z here. This is most conveniently

accomplished as a series of examples.

Case 1. Circular BeamUniform Circular Beam Center

Say that the illuminating intensity is Gaussian

ut(-P,), 2=t exp(- p'

TrTrb(4. F. 11)

where the beam center

(x,y) (4.F.12)

is random and circularly distributed

(4.F.9)

Page 74: ATMOSPHERIC PROPAGATION

-74-

(X-m )2+(y -M my)2 < R'7TR2

p (XY)

0

(4.F.13)

elsewhere

An unresolved target of area AT is assumed to be located at P' =

so that the reflected power is

A PP AT ut() 2 t t exp(- 2/r2)

r2 mTrb

(4.F.14)

Since pm is random it is clear that the reflected power is also. For

lijil = (m + m2) < R of this on = <IxI>z is given by the probability

density

r 2~2 b 1 CosTr R2 Z

-2rb ln(Z/Zmx) - R2 + I

21r1 rb/ nZ/Zmax IZmin ZZ mid

Zmid <Z <Zmax

2

2

0 el sewhere

(4.F.15)

where

PZ (Z) =

Page 75: ATMOSPHERIC PROPAGATION

-75-

e1

2 du u exp - u2 R0 2

.0 b b

Zmid exp

Zm =exp

1 - R)2

b

... R ( )2

b

Zmax

max

In (4.F.16),

first kind.

Io(-) is the zero-order modified Bessel function of the

For Jml > R the PDF of z becomes

2 2 1 2 r ln(Z/Zmx R2 + |i2

2 b 1 cos )T R 2 Z 2m1 rb/- 2 ln(Z/Zmax)

pZ (L) = .

Z* <_< Zmid(m4i nF.9)

(4. F. 19)

el sewhere

Zmax (4.F.16)

(4. F. 17)

(4.F.18)

m12 /2r2

0

Page 76: ATMOSPHERIC PROPAGATION

-76-

where Zmax' Zmid' and Zmn are again given by (4.F.16)-(4.F.18). A

sketch of this density is shown in Figure 4.3 for lin/rb = 0 and

R/rb = 1. The saturation SNR for this fluctuation with -mi/rb = 0 is

is given by

SNRSAT (4.F.20)bw R2 1 - exp(-2R 2/r )

2r (1 - exp(-R 2 /r ))2

and is sketched versus R2 /r in Figure 4.4.

Case 2. Fan Beam, Uniformly Distributed Beam Center

The M.I.T. Lincoln Laboratory mobile infrared radar is

sometimes operated with a beam shape that is not circular as in case

1, but is instead expanded in the vertical direction to produce a fan

beam [6]. Furthermore, when in scanning mode the vertical scan is

required to move only 1/128th as fast as the horizontal scan and

therefore it appears that the vertical aiming error is negligible.

With this in mind the target illuminating intensity is assumed to be

P- 2 exp(-(x' - x)2/r2) (4.F.21)

where Py has units (W/m) and the beam center x0 is random with

distribution

Page 77: ATMOSPHERIC PROPAGATION

W W W w

N

N0~

0.6 0.7 08 0.9 1.0 1.1 1.2 1.3z

Figure 4.3: Target return PDF for beam wander fluctuations; Gaussian beam,

uniform circular beam center, tII/rb = 0.0, R/rb =l..

01 I I I I I

2-

1

-4-4

116597-N

qW MW 1W W

Page 78: ATMOSPHERIC PROPAGATION

qwW

I 16(b 1-N

I I I I I I I I I

50 GAUSSIAN BEAM,CIRCULARLY DISTRIBUTED BEAM CENTER

40 --- FAN BEAM,UNIFORMLY DISTRIBUTED BEAM CENTER

2 3000

Z 20

10 - -

0

0 1 2 3 4 5 6 7 8 9 10

Saturation SNR due to two types of beam wander fluctuation.

low

Figure 4.4

Page 79: ATMOSPHERIC PROPAGATION

-79-

1

p (X) =

0

m - R < X < m+ R

(4. F.22)

el sewhere

such a uniform distribution is reasonable if one considers the

friction in the bearings that hold the beam aiming mirrors as the

dominant influence on the aiming error. Given (4.F.21), (4.F.22) and

an unresolved target located at p' = 0, the probability density for z

becomes for 0 < m < R

rb/2R

Z[-2 ln(Z/Zmax

rb/ R

z[-2 ln(Z/Z )]2

,max

0

Z <_ Z <_ Zmid

Zmid -Z < Zmax

el sewhere

(4.F.23)

where

PZ (Z) = .

Page 80: ATMOSPHERIC PROPAGATION

-80-

1

-r

2 R 0 rI b_

(4.F.24)

Q m+ R

b

Zmid = exp

min =

Q(y) =

y /i

and

1

1

(m - R)2r2rb

(m + R)2

rb

Zmax

Zmax

exp(-x 2 /2) dx

is related to the complementary error function by Q(y) = erfc(y/42)/2.

For m > R the pdf for z is

r b /2R

Z[-2 ln(Z/Z max 2

pZ (Z) =

Zmin Z < Zmid

(4.F.28)

elsewhere

Zmax

(4.F.25)

(4.F.26)

(4.F.27)

0

Page 81: ATMOSPHERIC PROPAGATION

-81-

where Z max Zmid, Z are again given by (4.F.24)-(4.F.26). The

saturation SNR for this case is given by

SNRSA = R1(4.F.29)SSATbw r Q{2(m-R)/rb} - Q{2(m+R)/rb -1

7 rb [Q{v2 (m-R)/rb} - Q{W2 (m+R)/rb1 2

A sketch of pz (Z) is shown in Figure 4.5 for m/rb = .45 and R/rb = 1.2.

It is interesting to note that the step in pz (Z) is due to the fact

that the reflected power depends not on the random variable x0 , but on

|x0 |. SNR is shown as a function of R2/rl for m/rb = 0 in FigureSATbb

4.4 and as a function of R/rb for various values of m/rb in Figure 4.6.

Case 3. Fan Beam, Gaussian Distributed Beam Center

Here we again assume the fan beam (4.F.21) but assume that

the beam center x0 is Gaussian distributed

p ( = 1 exp[-(X - m)2/2cy ] (4.F.30)0 / 2T2

x

Such a distribution for x0 is reasonable if the dominant influence on

beam wander is a thermal noise voltage in the radar aiming mechanism

or tubulent atmosphere beam steering [1]. For the case of an

unresolved target located at P' = 0 the probability density for z is

Page 82: ATMOSPHERIC PROPAGATION

T.5 I I I j I I I I I I I I I I I I I -

5.0-

2.5-

0.5 0.75 1.0 1.25z

Figure 4.5: Target return PDF for beam wander fluctuations; Fan beam, uniformly

distributed beam center, m/rb = .45, R/rb = 1.2. 116600-N

Page 83: ATMOSPHERIC PROPAGATION

60 1 1 I 1 1 1 1

50- m/rb 0

40-

~in~30

Z 20

10- 0.8

0 - .

-10 1 1 1 1 I I I0 12 3 4 5

R/rb

116594-NFigure 4.6: Saturation SNR due to fan beam, uniform beam center

fluctuations.

Page 84: ATMOSPHERIC PROPAGATION

-84-

-2rb x Z7r b/aX az IFx

[-2 ln(Z/Zmax)] ZJ

exp(-m 2 /2a ) mrb[-2 ln(Z/Zmax -2

PZ(ZZma cosh b CmaxF

0< Z < max

0 elsewhere

(4.F.31)

where Zmax is given by

m 2/2r2+ = (1 + a/r )2 exp b (4.F.32)

max x + CF2/r2

A sketch of (4.F.31) is shown in Figure 4.7 for m/rb = .8 and

a /rb = 1.25.

Page 85: ATMOSPHERIC PROPAGATION

4 III Ij III 111111 1111111 liii liii 111111 liii

N

N

0 0.25 0.5 0.75 1.0 1.25 1.5 1.75 2.0

z

Figure 4.7: Target return PDF for beam wander fluctuations; Fan beam, Gaussian 116598-Ndistributed beam center, m/rb = 0.8, a X/rb = 1.25.

Page 86: ATMOSPHERIC PROPAGATION

-86-

Case 4. Circular Beam, Gaussian Circular Beam Center

Here we again assume that Gaussian intensity (4.F.ll) but

assume that the beam center P = (x,y) is Gaussian distributed

p (XY) - 2 exp{{(X-m )2 + (Y-m )2]/2a2} (4.F.33)xy2'rr 2 X Y

This distribution on pm is reasonable if beam wander is dominated by

thermal noise in the radar aiming mechanism or turbulent beam

steering. For an unresolved target located at P' = 0 the probability

density for z is

r21 rb expV _ i

Zmax c 2 2

rb

pz(Z) =- Ia 2 /- bn /max 0 L m

o <7Z<7Z- -- max

0 elsewhere

(4.F.34)

where

Page 87: ATMOSPHERIC PROPAGATION

-87-

m12 /r 2

Zm = [1 + a2 /r ] exp 1b (4.F.35)zmax /b]exP2 1 + -a2/ r21 ~ bj

It is clear that beam wander can severely limit SNRSAT. It

is of interest to note (from Figures 4.4 and 4.6) that the limitation

is not too severe so long as the beam center varies within the area

of a diffraction limited spot (i.e. R/rb < 1) but becomes severe when

the opposite is true (R/rb > 1).

G. Correlation of Simultaneous Speckle Target Returns

In the introduction it was mentioned that the Lincoln

Laboratory compact CO2 laser radar employs a one-dimensional twelve-

element detector array. The transmitted energy is matched to this

array in that it is a fan beam, as mentioned earlier, compressed in

one transverse direction and expanded in the other. This situation

could be thought of as the detection of twelve simultaneously

transmitted, coherent, laser pulses. In order to make use of these

simultaneous measurements it is of interest to consider the

correlation between them. In the regime where the noise n is

negligible we have from (4.A.2) that the output of any matched-filter

envelope-detector is Ir 12 =XT) 2 where x(T) is given in (4.B.1).

Page 88: ATMOSPHERIC PROPAGATION

-88-

Hence, we would like to determine Cov( x1 )I 2 1 T2)[), the

covariance between the simultaneous target returns from two directions.

This is a difficult calculation to make directly. But, if x( T '

x(fT2) are zero-mean, complex Gaussian random variables, as is the case

for a large speckle target, and < x(l ) T2 )> = 0, which is true when

we can say <h L(P' 1 ) hL P 2)> = 0, then

Cov(Ix(Y ) 2I(T 2 Tl T2)> 2 (4.G.1)

so that we now need only the correlation function <x x*>. From this

quantity, the correlation angle, ec, will be found, where

<x(Tl) X*T2)> 0 for the arrival angle difference I fTl - fT2I Z 'c'

For the Gaussian spatial modes (4.B.2) and (4.B.3) with

PT = PR, the "effectively bistatic" assumption, the MFS propagation

model with poR = PoF = po and a speckle target, it is a tedious but

straightforward calculation to show that

- ntpPT - 2 'L X2_

<X( eTTa 2 exp [-IXLd 12/Por0 Trrb

{ dp' Ts(p') exp[-Ip' - Lfc 2/r]2 exp j2r r d-(' - XLTcj

S rres

(4.G.2)

Page 89: ATMOSPHERIC PROPAGATION

-89-

where

fd fTl ~ T2 (4. G. 3)

f T +T (4. G. 4)

and the correlation distance pcor is given by

r 2

P2 b (4.G.5)c o r + T2-+_ (PT/Po) 2

From (4.G.2), (4.G.3) it is clear that

c cor/L (4.G.6)

We find that ec = XS NT in the two limiting cases PT o and

PT p o under far-field propagation conditions. It can be concluded

that the correlation angle eC is essentially independent of

atmospheric conditions and is given by the field of view of a single

detector. This implies that matched-filter envelope-detector outputs

from adjacent pixels can be averaged for improved SNR if the target

is speckle and larger than a diffraction-limited FOV. Also, for the

multipulse detection problem discussed in Chapter VI, adjacent pixels

of large speckle targets can be taken to be independent observations.

Page 90: ATMOSPHERIC PROPAGATION

-90-

H. Backscatter

As discussed earlier, when the monostatic radar is used in

inclement weather there may be significant backscatter return from

the hydrometeors and aerosols present in the propagation path. In

this section, the backscatter contribution to the matched-filter

envelope-detector output is found by considering first order multiple

scatter [28] as the propagation mechanism. The development begins by

characterizing a thin layer of scatterers as a planar target. The IF

signal due to a transmitted impulse of energy and this single-scattering

layer is integrated to give the impulse response. This is then

convolved with a rectangular power waveform (Figure 1.2) resulting in

the backscatter contribution to the matched-filter envelope-detector

output.

Consider Figure 4.8. The field ut(P',z) is incident on a thin

layer of scatterers of cross sectional area A centered at p' = p0.

The backscattered field u(p',z) is given by

Ur(I',z) = ut(',z) Tb(p',z) (4.H.1)

as in equation (3.A.3). We assume that T (',z) is a random quantity

and can be statistically characterized as thought it were a speckle

target

Page 91: ATMOSPHERIC PROPAGATION

w w qw

.9!t IZI)

!--r(I',Z)

14 z JKAz

Single scattering layer; -

backscattered field.

(P',z) is incident field and ur(P',z) is

RADAR

Figure 4.8:

) ' Az)

SCATTERING SLAB HASAPRA A ANDir IS

CENTERED AT D' = o

116601-N

w

Page 92: ATMOSPHERIC PROPAGATION

-92-

<Tb(p, Z) j(3 ,z)> = A2 Tb(i ,z) 6(_ - T (4.H.2)

Since each scattering particle is at a random position within this

layer (4.H.2) is reasonable. We seek to express Tb in terms of the

backscattering cross section ab which is normally considered [28].

The quantity ab associated with a given particle is the area

intercepting the incident radiation which, when scattered isotropically

into 47r steradians solid angle, produces an echo at the radar equal

to that from the particle. If the region between the scattering layer

and the radar is free-space and the radar is in the far field of the

scatterers

I = 4bTZ 2 |t( 0,z)|2 pA Az (4.H.3)

is the power density (j;/m 2 ) of the backscattered field on the z = 0

plane where p is the particle number density (m- 3). In terms of the

target model (4.H.1) this same density is given by

I = <J dp' ho(p,p') _t(p',Z) Tb (',z)2>(

(4.H.4)

u-t( P 0,Z)12 T b A

Z2

Page 93: ATMOSPHERIC PROPAGATION

-93-

Combining (4.H.3), (4.H.4) we have

T =ab(z) p(z)b4r Az(4.H.5)

where ab and p have been allowed to be functions of z.

Consider again Figure 4.8. The contribution to the IF complex

envelope from the scattering layer is from (l.A.14)

y(t) = 2 s-T(t - 2z/c) dp' Tb(p',z) -F( 9',z) ER(p',z)

(4.H.6)

If we let

PT(t) = |ST(t)|2 = E 6(t)T0 (4.H.7)

where E = 1, then by the theory of the first-order multiple scatter [28]

the mean square contribution to the IF complex envelope becomes

<1y(t)12> = 6(t - 2z/c) - ab(Z) p(z) dp'IE0(-1',z)12 IO(5,z)I2hv 0 47r {exp(-2 at(s) p(s) ds) Az

-0

(4. H. 8)

Page 94: ATMOSPHERIC PROPAGATION

-94-

where use has been made of (4.H.5), R F are free-space field

patterns and at is the single particle total cross section. Integrating

(4.H.8) over z, the mean-square IF impulse response is

<ly(t) 12> = dz 6(t - 2z/c) h1 tP 2ab(z) p(z) J dp'l (p',z)2j F(p ,z)|2

0 0

-z

-exp(-2 a t(s) p(s) ds)

-O(4.H.9)

Convolving (4.H.9) with the waveform

t T

P T(t) =

0

0 < t < t

elsewhere

gives the mean-square output of H(f) (Figure 1.3)

c2 t

< x~)2> =

(t-t )2 p

t pP T 2 I'dz hv0 4T b(z) p(z) d'1 R(p',z)|2|((p',z)2

z

exp(-2 J at(s) p(s) ds)

-0(4.H.l1)

(4.H.10)

Page 95: ATMOSPHERIC PROPAGATION

-95-

which is essentially the output of the matched-filter envelope-detector.

For the Gaussian spatial modes (4,B.2), (4.B.3) with PT =PR and a

uniform scattering profile ab(z) = ab' at(z) = at, p(z) = p (4.H.11)

becomes

ct

< 1 t1 nt p P T2 2 exp(-2atpz) dz<|x (tb2 h b eld

o 22)l+

[27rP T Z

(4.H.12)

Equations (4.H.11), (4.H.12) give the backscatter power incident on

the radar receiver normalized by the local oscillator shot noise

power and, as such, should be compared to the CNR. When this

normalized backscatter power is smaller than the CNR it can be

neglected. But, when it is larger it can dominate target return.

This will be considered in more detail in Chapter VII.

Page 96: ATMOSPHERIC PROPAGATION

-96-

CHAPTER V

THEORY VERIFICATION

The IRAR project's compact CO2 laser radar provides a good

opportunity for verfication of the preceding theory. The MFS theory

treats only scattered light. However, at the CO2 wavelength of 10.6 im

the albedo of a single scattering particle is rarely larger than a

half [37], so the extinguished free-space portion of the beam will

dominate the scattered portion in inclement weather. This implies that

verification of the turbid atmosphere theory is difficult, if not

impossible, with the IRAR system. Hence, our efforts were concentrated

on verifying the turbulence models, of Chapter IV, Section D and the

beam wander models of Chapter IV, Section F. In this chapter we begin

with a brief description of the CO2 laser radar used to make the

measurements. This is followed by a description of the techniques used

in the data analysis. In the final two sections the beam wander and

turbulence models are examined in terms of measured data.

A. Laser Radar Description

The compact laser radar system [6] employs a one-dimensional,

twelve-element HgCdTe detector, a transmit/receive telescope of 13 cm

aperture, and a 10 W, 10.6 -pm, CO2 laser, which is operated in pulsed

mode. The radar system can be operated in three modes: (1) full

scanning mode, (2) reduced scanning mode, and (3) staring mode. When

Page 97: ATMOSPHERIC PROPAGATION

-97-

the radar is operated in full scanning mode there are two frame

rate/imaging options. In the first option the full, vertically stacked,

twelve element detector array is employed along with a fan shaped

illuminating beam, that has been compressed horizontally and expanded

vertically to match the detector array shape. Each detector is

separated by 200 rads from its neighbors directly above and below in the

direction it views. A 60 x 128 pixel image is constructed by sweeping

the fan beam horizontally through 128 - 100 prad steps five times. This

image, which takes 1/15th seconds to produce, then fills a 12 mrad x 12

mrad FOV.

In the second full scan, frame rate/imaging option a single

detector is employed along with a circular-symmetry Gaussian shaped

illuminating beam. A 60 x 128 pixel image is constructed by sweeping

the circular beam through 128 - 100 prad steps 60 times where each of

the 60 horizontal sweeps is separated by 200 prads in direction from

the one immediately preceding it. This image which takes 12/15th

seconds to produce, again fills a 12 mrad x 12 mrad FOV.

In reduced scan mode, a single detector is employed along

with a circular-symmetry, Gaussian shaped illuminating beam. A

60 x 128 pixel image is constructed in the same way as in the second

full scan, frame rate/imaging option except that the horizontal and

vertical steps are one twelfth as large. The image, which takes

12/15th seconds to produce,fills a 1 mrad x 1 mrad FOV.

When the radar is operated in staring mode a single detector

is employed along with a circular-symmetry, Gaussian shaped illuminating

Page 98: ATMOSPHERIC PROPAGATION

-98-

beam. Radar returns are measured from a single diffraction limited

FOV of approximately 50 virads (diameter of the e~ contour).

Successive laser illuminating pulses are separated in time by

104 bpsecs.

In all modes of operation the IF portion of the heterodyne

detected photocurrents are linearly passband filtered and then video

detected. The output of the video detector is proportional to the

magnitude of the envelope of the input (making it proportional to the

square root of the target return power). The video detector output is

then digitally peak-detected to yield 8 bit range and intensity values

which are stored on magnetic tape for off-line processing. In the

limit of a large target return (i.e. CNR 10) a stored intensity value

is essentially the output of a matched filter envelope detector and,

in terms of our previous radar model, is proportional to lxi (Equation

(4.A.3)).

All the data that has been examined comprises intensity returns

from three targets: (1) A retroreflector of approximately 2 cm diameter

which is well modeled as a pure glint target; (2) a polished sphere of

approximately 10 cm diameter, which, again, can be modeled as a glint

target; and (3) a 1 m x 1 m flame-sprayed aluminum plate which can

be taken to be a pure speckle target. All data discussed in this

chapter was recorded under the condition of high CNR. We have used

this data to verify the models of Chapter IV, Sections D and F.

Specifically, efforts have been aimed at statistically verifying

Page 99: ATMOSPHERIC PROPAGATION

-99-

equations (4.D.4), (4.D.6), (4.D.9), (4.D.11), (4.F.15), (4.F.19),

(4.F.23), (4.F.28), and (4.F.29). In the next section, the data

analysis techniques used for this will be described.

B. Data Analysis Techniques

Two principal types of calculations were used in the data

analysis. The first is an estimate of the saturation signal-to-noise

ratio, and the second is a chi-square goodness of fit test against

theoretical probability distributions.

If the stored intensity values, denoted 1xil, i = i,...,N,

are independent and identically distributed the saturation SNR which

is defined in (4.A.10) as

SNR< _X12>2SAT :Var( x1 2 )

can be estimated by

m2

SN SAT =

where m is the sample mean of the squared data

Nm = xi |i2

(5.B.1)

(5.B.2)

(5.B.3)

and y 2 is the sample variance

Page 100: ATMOSPHERIC PROPAGATION

-100-

Na2 (jx2 - m)2 (5.B.4)Ni=1 -

The performance of (5.B.2) in estimating SNRSAT is considered to be

approximated by the performance of (5.B.3) in estimating m. The mean

and variance of (5.B.3) are

<M> =< (5.B.5)

Var(i) = Var(X12) (5.B.6)N

The ratio

= N SNR (5.B.7)Var(m) SAT

indicates that the standard deviation of the estimate m is l/(N SNRSAT)P

of its mean so that if we required, for example

(Var(m))2 < 100 (5.B.8)

for a ±1% RMS error we must have enough samples to satisfy

N SNRSAT > 1002 (5.B.9)

For the case of a speckle target, SNRSAT < 1 from which it follows

that a minimum of 104 samples are required for a ±1% RMS error. For

Page 101: ATMOSPHERIC PROPAGATION

-101-

a speckle plate it is not too difficult to record this many measurements.

Unfortunately, over the period of time required to record this many

samples the characteristics of the radar itself may change. For example,

when observing an unresolved target the mean beam center m (see

Chapter IV, Section F) in fact drifts in time so that, as will be seen,

only a maximum of a few hundred points are recorded under the same

conditions. More will be said about this stationarity issue later.

Suppose that we have k mutually exclusive, collectively

exhaustive outcomes for some experiment with theoretical probabilitiesk

of occurrence p1p2 k .Z pi = 1. In a chi-square goodness ofi =1

f-it test [38,39] the hypothesis that the outcomes of the experiment

are governed by this distribution is tested against the hypothesis

that they are not. If n independent trials of the experiment are run

we calculate

k (f. - np.)z2 np (5.B.10)

i=l

where f. is the number of occurrences of the ith outcome and

kSf. = n (5.B.11)

i=l 1

Denoting the calculated value of X2 as X we see that X = 0 indicates00

perfect correspondence with theory while X1 large tends to discredit

the hypothesis. A quantitative measure of the validity of the

Page 102: ATMOSPHERIC PROPAGATION

-102-

theoretical distribution P1 SP2 '3'' 'Pk is provided by the level of

significance a

a = Prob(x 2 > X 2 ) (5.B.12)0

where the above probability is calculated assuming the theoretical

distribution is correct. It can be shown [38] that the random variable

X2 (5.B.10) is approximately a chi-square random variable of k - 1 - m

degrees of freedom so long as npi > 5 for all i = 1,2,...,k and m is

the number of parameters in the theoretical distribution that are

estimated from the data. Generally a value of a areater than or equal

to .05 is regarded as verifying the theoretical distribution.

In the applications of this test that are found here, the

underlying probability distribution is that of a continuous random

variable. In this case, the pi, i = 1,2,...,k are calculated as the

probabilities of the outcome falling into one of k contiguous intervals.

C. Beam Wander

In this section, we compare radar data from the retroreflector

target at 1 km range to the predictions of (4.F.15), (4.F.19), (4.F.23),

(4.F.28) and (4.F.29) via five examples. The data for the first four

of these examples was taken while the radar was operating in full

scanning mode with the first frame rate/image option so that the fan

beam, uniformly distributed beam center model of (4.F.21), (4.F.22)

Page 103: ATMOSPHERIC PROPAGATION

-103-

is appropriate. The data for the last example was taken while the

radar was operating in full scanning mode but with the second frame

rate/image option so that the circular beam, uniform circular beam

center model (4.F.ll), (4.F.13) applies.

Typically, when the IRAR system observes an unresolved glint

target in full scanning mode, the fluctuations on the target return

are dominated by random aiming error. In fact, not a single example of

atmospheric fluctuations dominating aiming error fluctuations could be

found in all of the full scanning mode data processed. The examples

of this section then represent typical samples of full scanning mode,

unresolved glint target data.

At the same time as the IRAR data was being collected one-way

scintillation measurements, whose purpose was to provide an accurate

estimate of the state of the atmosphere, were also being made. This

setup consisted of a GaAs laser and CO2 laser located approximately 10 m

to the side of the targets, and sensors corresponding to these lasers

located 10 m to the side of the radar. This equipment then provided

values of turbulence strength C2 and log-amplitude perturbation variancen

a2 at two wavelengths. For the first four examples of this section theX

scintillation measurements indicate that a2 was approximately 0.0005.X

From (4.D.6), (4.F.6)

SNRSAT = SNR 125 (5.C.1)STgu SAT ta

Page 104: ATMOSPHERIC PROPAGATION

-104-

The estimates of the saturation SNR, IEhSAT, for these examples were

much smaller than (5.C.1) so that we conclude (4.F.4), (4.F.5) are

satisfied and the beam wander fluctuations dominate target/atmospheric

fluctuations. The scintillation measurements indicate that at the

time the data for the fifth example was taken cy2 = 0.0026 so thatX

SNRSAT = SNRSAT ~: 24 (5.C.2)

gu ta

Again T was much smaller than this value and we conclude thatSAT

beam wander fluctuations dominate target/atmospheric fluctuations.

Figures 5.1-5.5 summarize the results of these five examples.

Figure 5.1 shows the theoretical distribution (4.F.23) with R/rb = 1.3,

m/rb = 0.0 along with a normalized histogram of 100 consecutive data

points taken over a period of approximately 7 seconds. The values

of R/rb and m/rb given above for the theoretical PDF were chosen to

minimize the calculated value of chi-squared as are all parameters of

theoretical PDF's in all examples of this type in this chapter. This

minimized, calculated value of chi-squared for this data is X = 4.10

with 8 degrees of freedom. This corresponds to a level of significance

a. between .8 and .9 indicating excellent agreement between theory and

data. For R/rb = 1.3, m/rb = 0, equation (4.F.29) gives SNRSAT = 8 dB

while from the data and (5.B.2), I'SAT = 6.6 dB, a difference of only

1.4 dB. Figure 5.2 shows the distribution (4.F.23) with R/rb = 1.2,

m/rb = 0.50, and a normalized histogram of 200 consecutive data points

Page 105: ATMOSPHERIC PROPAGATION

Z (Z)

6.0

5..

4.0

3.0

a.0

I.*

0.0#.So

ruiui~v u v ; i

I I

p

I I

0.60 0.70 0.30 0.90 1.06 1.10 1.10 1.30 1.40

Z

Figure 5.1: Normalized histogram of 100 consecutive retro returns taken in full scanning

mode and theoretical PDF Eq. (4.F.23), R/rb = 1.3, m/rb = 0.0.

I

. .I .

0

i

Page 106: ATMOSPHERIC PROPAGATION

t I I IF I I -F- I I I I I I I I I I I I I I

S. 9 I-

pz(Z)a.s L.

e.g I I II I I I i lA

0.50

4

I-. md - - m~~ ~ -

0.75 1.90

I I I

LL..L1.25

Z

Figure 5.2: Normalized histogram of 200 consecutive retro returns taken in full scanning

mode and theoretical PDF Eq. (4.F.23), R/rb = 1.2, m/rb = 0.5.

__j

I

Page 107: ATMOSPHERIC PROPAGATION

6.0

5..

4.0

3.0

e.9

i.e

0.

r Fi-l P El 1111 1111 1I IF 1

,.. E

S.as e.5. 0.75

I I II."

--

1.2s

I,

I 4

1.s0

Z

Figure 5,3: Normalized histogram of 200 consecutive retro returns taken in full scanning

mode and theoretical PDF (4.F.23), R/rb = 1,3, m/rb = 0.56.

M. - -

I

II I

Page 108: ATMOSPHERIC PROPAGATION

-108-

taken over a period of approximately 14 seconds and begun approximately

47 seconds after the finish of the data of Figure 5.1. Chi-squared was

calculated to be X2 = 14.54 with 12 degrees of freedom corresponding0

to a level of significance a between .25 and .30, still an excellent fit.

For R/rb = 1.2, m/rb = 0.50, equation (4.F.29) gives SNRSAT = 5 dB while

(5.B.2) gives $ITSAT = 5.9 dB. In Figure 5.3, R/rb = 1.3 and m/rb = .56.

The histogram is of 200 data points taken over 14 seconds beginning

20 seconds after the finish of the data of Figure 5.2. Chi-squared was

calculated to be X = 19.17 with 13 degrees of freedom corresponding to

a level of significance a between .10 and .20 which is still quite

acceptable. For R/rb = 1.3 and m/rb = .56, SNRSAT = 4.5 dB, while

SNRSAT = 4.7 dB.

For the above three examples m/rb changed from 0 to .50,

47 seconds later, and then to .56 another 20 seconds later. This gross

aiming error change of approximately one half a diffraction limited

FOV could be due to many causes. Movement of the people inside the

mobile radar vehicle or a change in the wind velocity against the side

of this vehicle could easily be responsible. This difficulty in

dealing with drifts and non-stationary effects is typical of highly

quantitative atmospheric propagation experiments.

The radar return corresponding to the pixel immediately to

the side (right or left, depending on which side of the retro the beam

center is) of the pixel in which the retro return is strongest should

also be dominated by the retro return. If the beam wander model

Page 109: ATMOSPHERIC PROPAGATION

-109-

(4.F.21), (4.F.22) correctly describes the situation then the target

return corresponding to this "side" pixel should be distributed

according to (4.F.28) with the same R/rb as the neighboring "hot"

pixel. Further, since there are two diffraction limited F0V's (e

points) between beam centers (which are separated by 100 yirads), m/rb

from the "hot" pixel plus m/rb from the "side" pixel should sum to 2.

In Figure 5.4 is shown a normalized histogram of 200 consecutive

points from such a "side" pixel along with the PDF (4.F.28) with

R/rb = 1.2 and m/rb = 1.43. The 200 data points of this figure

correspond to the "side" pixel of the 200 data points of Figure 5.2.

The value of R/rb which best fits the data is 1.2 in both cases and

the sum of the m/rb' s corresponding to these two pixels is 1.93.

Chi-squared was calculated, for Figure 5.4, to be X2 = 21.14 with0

12 degrees of freedom corresponding to a level of significance a

between .025 and .05, indicating fair agreement between theory and

data. For R/rb = 1.2 and m/rb = 1.43, SNRSAT = 0 dB, while

SAT = -1.4 dB. Further examples of this type were difficult to

obtain in our data set. Presumably this is because the beam is not

well described by a Gaussian form beyond the e-2 points.

Figure 5.5 shows the final example of this section. This time,

the data was taken while the radar was operating in full scanning mode

with the second frame rate/imaging option so that the beam wander

model (4.F.11), (4.F.13) applies. The figure shows the. theoretical

PDF (4.F.15) with R/rb = 1.4 and m/rb = 0.6 and a normalized histogram

Page 110: ATMOSPHERIC PROPAGATION

a.w0

1.s@

1.00

1.00

I ~I I I1.50

Z

Figure 5.4: Normalized histogram of 200 consecutive retro returns from the "side" pixel of

Figure 5.2 taken in full scanning mode and theoretical PDF (4.F.28), R/rb = 1.2,

m/rb = 1.43.

T

I I I

Pz(Z)

0.50

0.S 'Ii I6.60

Ii

2.50I 1 11 - - --

1w

~t - I I a I I I I I I I I I I I I I I I I I I

11111

CDI

Page 111: ATMOSPHERIC PROPAGATION

111111 III I I I lit I I III II I I I

.U£ L

S.69

I i Iliad A I I I i I I III I I. . A . . . . . . . .. I . . - .0.75 1.S0 I.25 1.6s

III

-J

1.75

Z

Figure 5.5: Normalized histogram of 400 consecutive retro returns taken in full scanning mode and

theoretical PDF (4.F.15), R/rb = 1.4, m/rb = 0.6.

*.75

*.s*9

Pz )

0.2S

0.25

Page 112: ATMOSPHERIC PROPAGATION

-112-

of 400 consecutive target returns taken over a 5 minute period. The

calculated value of chi-squared was X = 26.08 with 8 degrees of

freedom indicating a level of significance a of approximately 0.001.

This small value of a and lack of agreement between the theoretical

PDF and histogram shapes tend to discredit the model (4.F.ll), (4.F.13).

But before any conclusions are drawn it should be noted that the data

in this example was taken over a period of 5 minutes as opposed to a

maximum of 14 seconds in the case of the first four examples of this

section. This data collection interval difference was due to the factor

of 12 difference in frame rate between the two frame rate/imaging full

scanning mode options. There is then a much larger opportunity for

non-stationary and unidentified effects to cause radar return

fluctuations in the final example than in the first four. To test

the model (4.F.11), (4.F.13) it seems necessary to acquire data in a

much shorter period than has been done.

What these figures and discussion indicate is that even in the

absence of target and atmospheric fluctuations, random aiming errors

can cause fluctuations to be impressed upon the target return. These

fluctuations are severe when the aiming error standard deviation is

comparable to a diffraction limited field of view R/rb = 1. In

Figures 5.1-5.4, R/rb was consistently equal to 1.2 or 1.3. As this

data was taken over a period of 90 seconds, in scanning mode, the aiming

error of the IRAR system is a fact and quite severe.

The above remarks apply to unresolved targets. For

resolved targets the implications of beam wander are not as severe.

Page 113: ATMOSPHERIC PROPAGATION

-113-

Namely, the beam wander provides a mechanism for spatially averaging

several diffraction limited spots in a single pixel. As the target is

already resolved this is not too limiting and, as will be seen in the

next section, the wander increases the appropriateness of the speckle

target statistical model.

D. Turbulence

The data processing examples given in this section can be

broken into two groups. The first group involves radar returns from

the retro and sphere (glint targets) which are expected to be

distributed according to (4.D.4), while the second group involves radar

returns from the flame sprayed plate (speckle target) and should follow

(4.D.9).

All data in the first (glint target) group was collected while

the radar was operating in reduced scanning mode as this data could be

processed to eliminate the effects of beam wander, whether due to

atmospheric beam steering or radar aimpoint jitter. This was

accomplished by selecting only the single, maximum intensity value,

data point from each 60 x 128 pixel frame. This maximum point, termed

the "hot spot," should then be lognormally distributed from frame to

frame. In Figure 5.6, a normalized histogram of 100 consecutive. retro

reflector returns selected according to the above procedure is shown

along with the lognormal distribution

Page 114: ATMOSPHERIC PROPAGATION

3.,1

a..

Pz )

1.0

0.0SL I I I I I U I I I I U I U I I.o 0.4 0.3 1.3 1.6 3.0 1.4 3.3 2.3 3.1 4.0

-J

4.4

Z

Figure 5.6: Normalized histogram of 100 "hot spot" retro returns taken in reduced scanning mode

and the lognormal PDF (5.D.1), g 2 = .0045.x

Iw

Page 115: ATMOSPHERIC PROPAGATION

-115-

p (Z) = 1 exp- (ln Z + 2u 2) U(Z) (5.D.l)z /2 2a Z 18G2 X

X L X

with g 2 = 0.0045. Simultaneous scintillation measurements indicateX

that a2 = 0.0027, but, the value of 0.0045 was chosen to minimize theX

calculated value of chi-squared. This calculated value was X2 = 12.70

with 7 degrees of freedom indicating a level of significance a between

0.05 and 0.10 and fair agreement between theory and data.

In Figures 5.7 and 5.8, the time evolution of the location of

the hot spot within the 100 - 60 x 128 pixel images used for the

example of Figure 5.6 is shown. Specifically, Figure 5.7 shows the

row location with separation between rows corresponding to an angle

of 17 Vrads, while Figure 5.8 shows the column location with

separation between columns corresponding to an angle of 7.8 Prads.

These figures indicate that the RMS angle error of the radar return is

on the order of 15 prads. The independent scintillation measurements

indicate that the RMS angle error due solely to atmospheric beam wander

effects is X/p ~ 25 prads. Since these two numbers are very nearly

equal, and any aimpoint jitter would cause the RMS angle error of the

radar return to increase, we can conclude that the beam wander and

non-constant behavior of the hot spot location is due to atmospheric

beam steering effects. We can also conclude that while in reduced

scanning mode the frame-to-frame RMS aiming error is less than

Page 116: ATMOSPHERIC PROPAGATION

LU

3117 Brad

29

tThtYt-t-t-t-99999+999@~@-i--I-f** I40@+-4 O4* 14@4-14 [-14-1-14

0B 27

25

I FFVrFrrTIYrrmTYY1-l-Y-tIY1-t~tw-.- -t+9. l-. H- 1 4-01 1 -I-44 . &4-&-464 14-44--1 1-1 1 1 &

0 20 40 60 80 100

FRAME NUMBER

Figure 5.7: Time evolution; Row location of hot spot in the 100 - 60 x 128 pixel frames used in

the example of Figure 5.6.

( I I I I I I I I I I I I f I I I I I I I I I I I I i I I I I I I I I I I I I I I I r- . I - I . . I I . . I . , . . ,. . . . . . . . ..

-- --- --

I W I F-r-T--r-T

.W 1W

i ; i

I

I

I111MkIT1MU1,

Page 117: ATMOSPHERIC PROPAGATION

Mw w 1w

I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I T

84

7.8 prad82

80

C,

LYITI----HT±.---- ---- ----

TF~TTfl~l7MftTitu1 Ufi llLLTIUJJ tl1LtIJ~l~I1~itIU73

75C

11111120 40 80 10060

FRAME NUMBER

Figure 5.8: Time evolution; Column location of hot spot in the 100 - 60 x 128 pixel frames used in

the example of Figure 5.6.

-4

f I I I . i i i i 1 1 1 1 1 1 1 1 i

-" - . . . . 1 1 1 1 1 1 1 1 1 1 1 L--L

. . . . .. . . . . . . . I . I . . I . I . . I . ; i +-V4-t-4 -+-4-4-4-4-4M4-4 I 1 4

- - -- - - - - If I I I I I I I I I I I I I I I I I I I I I I1-

I

Page 118: ATMOSPHERIC PROPAGATION

-118-

25 yprads. This is in contrast to full scanning mode in which we saw

this RMS aiming error was approximately R/v'3 rb = .7 diffraction limited

F0V's or 35 yirads.

In Figure 5.9, a normalized histogram of 300 consecutive

"hot spot," retro reflector returns is shown along with the PDF (5.D.1)

with a2 = 0.0138 to minimize X2. Simultaneous scintillation measurementsX

indicate that a2 = .0189. Chi-squared was calculated to be X2 = 12.78X 0

with 9 degrees of freedom for a level of significance a between 0.10

and 0.20 indicating very good agreement between theory and data.

Figure 5.10 shows a normalized histogram of 400 consecutive

"hot spot" retro returns along with the PDF (5.D.1) where a'2 = 0.018X

to minimize X = 12.85 with 15 degrees of freedom. The level of

significance a for this example is between 0.50 and 0.70 indicating

excellent agreement between the data and theory. The scintillation

measurements give a2= 0.055, a factor of three larger than the X2X

minimizing value, though.

In Figures 5.11 and 5.12, histograms of 300 consecutive "hot

spot" polished sphere returns are shown along with the PDF (5.D.1).

In Figure 5.11, a2 = .0083 minimizes X = 10.19 with 12 degrees of

freedom indicating a level of significance a between 0.50 and 0.70 and

excellent agreement between theory and data. The scintillation

measurements give a2 = .0245 for this figure. In Figure 5.12,X

a2 = .004 minimizes X2 = 11.93 with 11 degrees of freedom indicatingXa level of significance a between .30 and .50 and, again, excellent

Page 119: ATMOSPHERIC PROPAGATION

w w

*.0 0.4 0.3 l.a 1.6 3.0 3.4 8.S 3.1 3.6 4.4

Z

Figure 5.9: Normalized histogram of 300 "hot spot" retro returns taken in reduced scanning mode

and the lognormal PDF (5.D.1), a2 = 0.0138.x

3.S@

1.50

i.4s

Mo,

pZ(Z)

SIII iI i 11 1 1 1111111113i -1

.mE.'-

4.4

qw

Page 120: ATMOSPHERIC PROPAGATION

77(

I I I I- - I * ~ I I U 1 I .d. I & 4

1.016

iii j~~l

1. 5 11.4

Z

Figure 5.10: Normalized histogram of 400 "hot spots" retro returns taken in reduced scanning mode

and the lognormal PDF (5.D.1), a2 = 0.018.x

-1I

1.50

1.86

P (Z)

0.60

I I III

6.8 O.6 0

M)

I

Page 121: ATMOSPHERIC PROPAGATION

MW RW

pz (Z)

0. -IL I I I I

tse

0.0SAS5 1.00 1.50 3*

pzz

Z

Figure 5.11: Normalized histogram of 300 "hot spot" polished sphere returns taken in reduced

scanning mode and the lognormal PDF (5.D.1), a = .0083.

Page 122: ATMOSPHERIC PROPAGATION

w Iw

3.,

2.9

p (Z)Z

1. .

S.. 10..

gI II li ii i i i li i i i I Il

I I I I I I . i I L I-i i ~ ~

0.S0 I.49 '.se a."

Z

Figure 5.12: Normalized histogram of 300 "hot spot" polished sphere returns taken in reduced

scanning mode and the lognormal PDF (5.D.1), cV2 = 0.004.x

. . . . .. .I _ . .I . . . . .1 1 1

Page 123: ATMOSPHERIC PROPAGATION

-123-

agreement between theory and data. The scintillation measurements give

2 .0026 for this figure.X

The single glint SNRSAT curve (4.D.6) is shown in Figure 5.13

along with the estimates fN SAT calculated from the data of Figures

5.6, 5.9-5.12. Each of the five values of SONSAT is plotted twice;

once vs. measured a2 values (from the scintillation measurements),X

indicated by squares, and second vs. the value of a2 that minimizesX

X", indicated by circles. Here we see very good agreement between the

predictions of (4.D.6) and the data, at least within the limited range

of available a2 values.X

The data in the examples of the second (speckle target) group

was collected in all three modes of operation. In Figure 5.14, a

normalized histogram of 2000 consecutive speckle plate, squared,

intensity returns taken by the radar operating in full scanning mode

with the first (fan beam) frame rate/imaging option is plotted along

with an exponential PDF. The data in this example (and only this

example) is squared. As simultaneous scintillation measurements give

a2 = .02, Equations (4.D.11) and (4.D.9) indicate that this squared

data should fit an exponential distribution. The calculated value of

X = 17.84 with 11 degrees of freedom indicating a is between 0.05

and 0.10 and reasonably good agreement between theory and data.

The data of Figure 5.15, which shows a normalized histogram of

400 speckle plate returns and a unit mean Rayleigh PDF

Page 124: ATMOSPHERIC PROPAGATION

-124-

10

2a

100

- X

102= o -points plotted against minimizin

2

1O6 2 ol pints lotted aga ins minimizig

10' 8"10 1

22

xX

100

10 4 1I3 f

Figure 5.13: The theoretical curve (4.D.6) and estimates 'SNASAT fromthe data of Figures 5.6, 5.9-5.12.

Page 125: ATMOSPHERIC PROPAGATION

I I I I I I 1 1 1 1 1 1 1 1 I0 . 0040

0 . 0030

0.0020

Pz (Z)

0.001 0

0 0- --- I I .L iJ I I I I

0. 500. 1000. 1500. 2000.

Z

Figure 5.14: Normalized histogram of 2000 squared, consecutive speckle plate returns taken

in full scanning mode and exponential PDF.

N,

Page 126: ATMOSPHERIC PROPAGATION

w Aw w

0.75

0.50

pz )

0.25

0.as

0 .0, 0.4 0.8 1.8 1.6 8.0 P.4 2.8 3.2 3.6 4.0 4.4

Z

Figure 5.15: Normalized histogram of 400 consecutive speckle plate returns taken in full

scanning mode and Rayleigh PDF.

K 1 7 =I11 , 11 1 1 1 1 1 1 1 1 1

1 1 I I I I I I 1I 11 I9 1 1 1-

a.)

Page 127: ATMOSPHERIC PROPAGATION

-127-

Pz(Z) = Z exp[-TrZ 2 /4] u(Z) (5.D.2)

was also collected in full scanning mode, as in the previous example,

but with the second (circular beam) frame rate/imaging option.

Scintillation measurements give a2 = .004 so that by (4.D.ll), (4.D.9)

the data should be distributed according to (5.D.2). The calculated

value of x2 = 14.92 with 16 degrees of freedom indicating a is between0

0.50 and 0.70 and excellent agreement between theory and data.

In Figure 5.16 a normalized histogram of 1200 data points is

shown along with the Rayleigh PDF (5.D.2). The 1200 data points were

taken from 48 consecutive 60 x 128 pixel reduced scan images. The

25 data points from each frame were arranged in a 5 x 5 matrix across

the target, each separated in angle by approximately 200 lirads from its

nearest neighbors. Scintillation measurements give a2 = 0.004 so that,

again, the data should be Rayleigh distributed. The calculated value

of X2 = 23.44 with 19 degrees of freedom indicating at is between 0.200

and 0.25 and very good agreement between theory and data.

In Figure 5.17, we have a normalized histogram of 1400 speckle

plate returns taken in staring mode along with the Rayleigh PDF

(5.D.2). Scintillation measurements give a2 = 0.004 at the time

these measurements were made so that the data, according to (4.D.9),

(4.D.11), should be Rayleigh distributed. But contrary to this the

data has a significantly narrower distribution than the theoretical

prediction. Further, the calculated value of chi-squared is X = 2850

Page 128: ATMOSPHERIC PROPAGATION

w

@.76

0.50

Pz )

.as

0.00.0 0.4 0.3 1.a 1.6 2.6 3.4 3.3 3.8 3.6 4.0 4.4

Z

Figure 5.16: Normalized histogram of 1200 speckle plate returns taken in reduced scanning mode

and Rayleigh PDF.

I I I U I A I I I iI I I

N)0,

low Vw

Page 129: ATMOSPHERIC PROPAGATION

w w w

N

i-i III I LI

T

11111111* 0.4 0.6 1.3 1.6 a.$ 3.4 i.E 3.8 3.6 4.0 4.4

Z

Figure 5.17: Normalized histogram of 1400 consecutive speckle plate returns taken in staring

mode and Rayleigh PDF.

1.00

0.5

pz(Z)

0.26

0.

IIIj

- I I I I I I I I I I I I I I I I A I I I I

_ _ _ _ _ _

Page 130: ATMOSPHERIC PROPAGATION

-130-

with 18 degrees of freedom indicating extremely poor correspondence

between theory and data. To understand this disparity between theory

and experiment we must consider the nature of a speckle target. If we

were to view a "rough" target through free space in starina mode and

both the target and radar were held perfectly fixed in space, i.e., all

movement and vibrations had been eliminated, then it is clear that all

target returns would be the same. That is, under the above conditions,

the target return probability distribution would be an impulse. It is

only when, due to atmospheric beam steering effects, radar aiming

errors, and target movement, we begin to sample several diffraction

limited spots on the target that we begin to see the Rayleigh

statistics. In Figure 5.17, we are at an intermediate stage between

looking at a single diffraction limited spot, resulting in an impulsive

data distribution, and sampling many spots, resulting in Rayleigh or

exponential statistics as in Figures 5.14-5.16. So that we conclude

that the speckle nature of a target depends on more factors than just

the roughness of the target surface and is closely related to how many

diffraction limited spots on the target are sampled by the radar.

The final example of this section is summarized in Figures

5.18 and 5.19. The same normalized histogram of 300 speckle returns

taken in full scanning mode with the second frame rate/image (circular

beam) option is shown along with a Rayleigh PDF in 5.18 and a Rayleigh

times lognormal PDF (see Equation (4,D.9) and [27]) in 5.19. The value

of X2 = 13.51 with 20 degrees of freedom in Figure 5.18 indicating a0

Page 131: ATMOSPHERIC PROPAGATION

1.80

0.75

0 * 50

Pz )

OS 0 0.4.- , -- I I I - I-A L dL .

Z

Figure 5.18: Normalized histogram of 300 speckle plate returns taken in full scanning mode

and Rayleigh PDF.

CA)

MW

1.8 8,.$ 3~ -. 8~ ~ 3.9 4.0 414

Page 132: ATMOSPHERIC PROPAGATION

qw MW

0.76

*aEO

p (Z)Z

0.26

Ss.60.0 0.4 0.1 1.a 1.8 3.0 3.4 eA. 3.8 .8 4.0 4.4 4A1

Z

Figure 5.19: Normalized histogram of the same 300 speckle target returns as Figure 5.18

and a Rayleigh times lognormal PDF, a2 = 0.01.

- -'

kil -L . .. I..

N)

Page 133: ATMOSPHERIC PROPAGATION

-133-

is between 0.80 and 0.90 and excellent agreement between the PDF and

the data. The calculated value of X2 is minimized by choosing 92= 0.010

to x' = 13.20 with 19 degrees of freedom in Figure 5.19 indicating a0

is again between 0.80 and 0.90 and excellent agreement between the

PDF and the data. For both figures, scintillation measurement gives

a2 = 0.15. At this level of turbulent fluctuations, a marked departure

from Rayleigh statistics is expected. But contrary to this, the data

fits a Rayleigh distribution very well and when fit to the Rayleigh

times lognormal PDF the X minimizing value of a2 is more than an order

of magnitude away from the measured value. Exactly why this is the

case is not as yet clear. But the scintillation measurements indicate

that the atmospheric coherence distance p is smaller than the radar

beam size. Hence, the model (4.D.9) does not strictly apply. It is

suspected that this fact plays a significant role in explaining the

divergence between theory and experiment here.

In this chapter, we have given many examples of our theory

verification efforts. The model for aimina error, at least in the case

of a fan illuminating beam, seems to be correct. Also, the lognormal

character of the atmospherically induced fluctuations, at least from

a glint target, appears to have been verified. What is less certain

is the Rayleigh times lognormal character of the speckle plate return

in clear weather. Indeed, the essentially free space result of

speckle plate Rayleigh statistics was verified. But in heavier

turbulence it was difficult to find examples which clearly followed

Page 134: ATMOSPHERIC PROPAGATION

-134-

(4.D.9). Exactly why this was the case is not clear. More effort

should be directed towards understanding this.

Page 135: ATMOSPHERIC PROPAGATION

-135-

CHAPTER VI

TARGET DETECTION

Here we discuss multipulse detection of targets at known

range by the radar. That is, we address the problem of optimally

deciding between the hypothesis that there is a target present at

range L and the hypothesis that there is no target present. This

decision is based upon the returns from M transmitted laser pulses.

The detection problem just posed is most relevant to objects viewed

from a ground-based radar against a nonreflecting background

(i.e. the sky) as background clutter is ignored. In the first

section of this chapter we formulate the problem mathematically.

Following this, expressions for single pulse performance are given.

Multipulse integration is discussed in the next section. Finally,

linear integration multipulse performance is presented.

A. Problem Formulation

Consider transmitting M laser pulses from the radar and

observing the IF returns ri(t), i = 1,2,...,M. These pulses can

occur sequentially as in Figure 1.2, concurrently but mutually

separated in angle by at least a diffraction limited FOV, or both.

If no target is present the IF signals are pure noise while if a

target is present the IF signals are return plus noise. Mathematically,

Page 136: ATMOSPHERIC PROPAGATION

-136-

if we call target absent hypothesis H0 and target present H1 the IF

complex envelopes corresponding to the M transmitted pulses satisfy

n (t) H 0

r.(t) = , (6.A.1)

-1 { :?:ng(t) 1

i = 1,2,... ,M

where r.(t) is observed from 2L/c seconds to 2L/c + t seconds after

transmission of pulse i. The xi are given by

'rtpP '2 r .i

x = hv) T df' T(p') ERip' E Fip' = 2ij e

(6.A.2)

where the 8. can be taken to be mutually independent, uniformly

distributed random variables over [0,27r]. The n (t) are mutually

independent, circulo-complex, white, zero mean Gaussian processes with

<n (t) nt(s)> = tp 6 6(t - s) (6.A.3)

As a scanning radar with frame rates slower than 20 Hz is assumed, the

IF returns r.(t), i = 1,2,...,M can be taken to be independent under

either hypothesis. The optimal Neyman-Pearson likelihood ratio test

(LRT) [23] for target detection is then

Page 137: ATMOSPHERIC PROPAGATION

-137-

H

L r1 (6.A.4)

H0

The likelihood ratio L is given by

M CX2L = II dX p e- I {2X (6.A.5)

where p, iI(X) is the probability density function for the magnitude

of the signal return x and

Ir.i = (1/tp) {r(t) dtj (6.A.6)

is the output of a matched filter envelope detector. The integral

(6.A.6) is taken over the t second interval beginning 2L/c seconds

after transmission of the ith pulse. In (6.A.5) I {-} is the

zeroth-order modified Bessel function.

As no assumption has been made regarding the statistics of

the xi this formulation applies to low visibility as well as turbulence.

In particular, if we consider the case of a single transmitted pulse,

M = 1, the LRT (6.A.4) reduces to the threshold test

H

2 < y (6.A.7)H0

Page 138: ATMOSPHERIC PROPAGATION

-138-

The optimality of the above test applies regardless of the target

scenario and atmospheric conditions. The performance of the test, as

we shall see in the next section, unfortunately is not similarly

independent of these details.

B. Single Pulse Performance

Here we discuss the performance of the Neyman-Pearson test

(6.A.7). This evaluation involves the determination of the

probability of detection PD and the false alarm probability PF

PD = Prob(1r_1 2 > yjH 1 ) (6.B.1)

PF = Prob( r12 > yjH0 ) (6.B.2)

where the subscript on r has been dropped.

The false alarm probability is independent of the

atmospheric/target scenario and depends only on the LRT threshold.

P = exp(-y) (6.B.3)

so that (6.A.7) becomes

H

Ir 2 - ln P (6.B.4)

H0

Page 139: ATMOSPHERIC PROPAGATION

-139-

In contrast, the detection probability depends on the atmospheric/target

scenario as well as threshold. We present our results as a series of

examples.

Case 1. jx2 Nonrandom

If rx! 2 = CNR is known, as in the case of a glint target in

free space, the detection probability is given by [3,23]

dz z exp(-(z 2 + 2!xI)_|2)I 0 (z LxI)D =12(-2lnP F) (6.B.5)

= Q(/2-- W, (-2 ln PF

Equation (6.B.5) is Marcum's Q-function.

Case 2. Single-Glint Target in Turbulence

For a single-glint target T(p') = eje T ('-_g(4.D.l)-(4.D.3) the detection probability is

PDdx p (x) Q[(2 CNRxA gue

) satisfying

[3,5]

X) e2X,(-21nPF) ]

where

(6.B.6)4A 2

Page 140: ATMOSPHERIC PROPAGATION

-140-

p(x) = exp[-(x + a2)2/2&21 (6.B.7)X X X

V2Tra 2

X

is the pdf for the log-amplitude perturbation X.

Case 3. Ix1 2 Exponential

Here we consider the case

IX2 = CNR v (6.B.8)

where v is a unit mean exponential random variable. Equation (6.B.8)

applies to the three scenarios:

1. Free space, speckle target.

2. Low visibility, large speckle target.

3. Low visibility, resolved glint target.

The detection probability is [23]

p = p(CNR + 1)l (6.B.9)D F

Case 4. Resolved Speckle Target in Turbulence

For a target whose radar return is described by (4.D.9) the

detection probability is [3,5]

Page 141: ATMOSPHERIC PROPAGATION

-141-

dU pu(U) P (CNRsre2U +l)_lu F

Pu (U) = exp[-(U + u2 )2 /2a2 )]

v/2Trcf2

is the pdf of the aperture-averaged log amplitude perturbation u.

Case 5. Ix12 A Product of Two Exponential Random Variables

The case of a small glint target in low visibility is

considered here

Ixi 2 = CNR v w (6.B.12)

where v, w are IID unit mean exponential random variables. The

probability density for the product y = vw is given by

py (Y) = 2K (2vY) u(Y)

where K (-) is the modified Bessel function of the second kind

order zero, and u(-) is the unit step function. A sketch of

(6.B.13) is shown in Figure 6.1. The detection probability in

case is given by the integral

(6.B.13)

of

this

PD

where

(6.B.10)

(6.B.11)

Page 142: ATMOSPHERIC PROPAGATION

-142-

2.0

1.6

... 1.2

0.8

0.4

0.0

0 1 2 3 4

Figure 6.1: The Probability Density Function, p (Y) = 2 K (2Y)yi 0

Page 143: ATMOSPHERIC PROPAGATION

-143-

( f in PP = dt exp -t + CNR t1 (6.B.13)

We are most interested in case 5 here as it represents new

work. Also, it is worthwhile, in a practical sense, to contrast this

case with case 2 as both (6.B.6) and (6.B.13) are detection

probabilities for small, single-glint targets in two limiting

atmospheres. In Figure 6.2 the detection probability PD is plotted

vs. CNR for Ix1 2 a product of two exponentials and various false

alarm probabilities. In Figure 6.3 we have the receiver operating

characteristics, PD vs. PF as the threshold y is varied, for several

values of CNR. From these figures we see that PD is a monotonically

increasing function of both CNR and PF. We also observe, from

Figure 6.2, that if just two or three dB of CNR can be obtained over,

say, the PF = 10~ case, the false alarm probability can be improved

by four orders of magnitude to PF = 1011 while maintaining constant

detection probability PD. The bumps found in the curves of Figure

6.2, at high CNR, are thought to be due to the numerical integration

of (6.B.13) and are not real oscillations. In Figure 6.4, PD is

plotted vs. CNR for |X 2 a product of two exponentials and a single

glint target in turbulence (Equation (6.B.6)) for PF = 10 7. Each

curve for a single glint target in turbulence corresponds to a

Page 144: ATMOSPHERIC PROPAGATION

--141-

99.99

99.9 SINGLE PULSE IXI2 =IX2>VW

99

95

90

2 70

P =1050 F

CL

30-

10

5 -pF =10-4-

P F P107

S1-11

0.1 F

0.01-20 0 20 40 60

CNR (dB)

Figure 6.2: Single pulse detection probability vs. CNR

for a glint target in bad weather.

Page 145: ATMOSPHERIC PROPAGATION

-145--

~_ SINGLE PULSE

1010

I I

X|2 2> VW

99.99

99.9

99

95

90

-6 106

Figure 6.3: Single pulse receiver operating characteristics for a glint

target in bad weather.

CNR 40 dB

_CNR=20 dB

CNR = OdB

70

50

30

C

C.,

a,0.

0C-

10

5

0.1

0.01-4 -2

I I

PF

Page 146: ATMOSPHERIC PROPAGATION

.l46-

99.99 1 1 1 1 / I 1 1

99.9 SINGLE PULSEP 0-7

F PF

99 2=

95- .2 = 0.01

x/x90-

2 = 0.1 2 =0.5

70-

2 2CL 50 - , |_ =< X( >VW

a. 30-

10

5-

1I

0.1 /0.01/

-20 0 20 40 60

CNR (dB)

Figure 6.4: Single pulse detection probability vs. CNR for a single

glint target in free-space, three levels of turbulent

fluctuations and bad weather. PF = 10~7 throughout.

Page 147: ATMOSPHERIC PROPAGATION

-147-

different value of a2 with a2 = 0 indicating free space. We observeX X

improved performance of the jxj 2 random curves over the jx1 2 nonrandom

curve (a2 = 0) in the region CNR 5 9 dB. This can be understood ifX

one considers the likelihood of a fluctuation bringing the matched

filter envelope detector output above threshold in this low CNR

regime. It is also evident from this figure that significant CNR

increases will be needed to maintain high PD values on a glint target

in the presence of strong turbulence or scattering. This figure

indicates that for equal CNR a glint target in the turbid atmosphere

is easier to detect than the same target for saturated scintillation

C2 = .5. The problem is that in turbulence the CNR is essentially the

free space CNR whereas, from (4.E.15), the CNR in bad weather can be

significantly reduced from the free space value. More will be said

about this in the next chapter. Again, as in Figure 6.2, the bumps

in the Ix1 2 = <tx1 2 > vw case are thought to be due to the numerical

techniques used to integrate (6.B.13).

C. Multipulse Integration

As should be evident from Figures 6.2-6.4 adequate detection

performance cannot be maintained at lower CNR for a single radar pulse.

In order to improve upon this situation we find it necessary to use

several pulses in combination. In this section, we address the

problem of approximating the multipulse decision rule (6.A.4) when

the signal return jx! 2 is a product of two exponential random variables.

Page 148: ATMOSPHERIC PROPAGATION

-148-

For CNR = CNR, i = 1,2,... ,M this is not too difficult. Consider

the likelihood ratio (6.A.5)

ML = i R (6.C.1)

i=1

where

R. = dX p (X) e-X IX{2X } (6.C.2)

For jx.1 2 = CNR v i wi, (6.C.2) becomes

R. = dX 2K 2 I{2X Ir 1} (6.C.3)S10 N jdX/CKo -NR j e o -I

This integral has been evaluated numerically using cautious adaptive

Romberg extrapolation [29]. The results are shown in Figure 6.5

where R is plotted vs. Ir[ for several values of CNR along with the

curve

log(R) = -3.10 + 0.344 Ir12 (6.C.4)

It is evident from comparing the plot of (6.C.4) and the CNR = 40 dB

plot of (6.C.3) that (6.C.3) is well approximated by a parabola

Page 149: ATMOSPHERIC PROPAGATION

-149-,

116765-M

0 1 2 3 4 5 6

1'r

Figure 6.5: Likelihood ratio R (Eq. (6.C.3)) and parabola

log R = 3.1 + 0.3441r|2 vs. matched filter

envelope detector output Irl.

10 8

106

105

10 4

R 103

10

10

10-2

10~

-3

-4101

101I

10

CNR = OdB

CNR 20 dB

log R = 3.1

+ 0.344 Irl2CNR zOdB

CNR 20 dB -

- CNR= 40dB -

log R 3.1 + 0.344 r|2

Page 150: ATMOSPHERIC PROPAGATION

-150-

log(R1 ) = A(CNR) + B(CNR) Ir 12

R. = eA(CNR) eB(CNR) r*iJ2

where we have explicitely

Using (6.C.6) in place of

gives, for equal CNR, the

Mi

noted that A and B are functions of CNR.

(6.C.3) in the likelihood ratio (6.C.1)

Neyman Pearson test

H

Ir 2 Y (6.C.7)

HO

This test is then very nearly optimal. The performance of (6.C.7) is

investigated in the next section.

D. Multipulse Performance

Here we consider the PD F behavior of the threshold test

Z= r

H

H0

Y

for several target/atmospheric scenarios. The use of (6.D.1) in

case of a small glint target in bad weather was justified in the

last section. If lxf 2 is an exponential random variable (6.D.1)

(6.D.1)

the

can

or

(6.C.5)

(6.C.6)

Page 151: ATMOSPHERIC PROPAGATION

-151-

be shown to be optimal for CNR = CNR, i = 1,2,...,M. At any rate,

use of (6.D.1) can be considered to be arbitrary if it is not optimal.

The false alarm probability PF of (6.D.1) is independent of

target/atmospheric scenario and depends only on number of pulses M and

threshold y. An exact expression is

p r(MMY) (6.D.2)F (M- 1)!''

where

F(a,x) = a- e-u du (6.D.3)

is the incomplete Gamma function [30]. Evidently exact performance

results are difficult to obtain for the test (6.D.1), and, once

obtained are generally cumbersome. Hence it is worthwhile developing

accurate approximate results. Towards this end, we use a modified

Chernoff bound procedure [31,32] to derive these results. This technique

is useful for estimating the area underneath the tails of a probability

distribution. Hence the results that follow only apply in the regime

PF 0.40, PD > 0.60.

Defining two conditional semi-invariant moment generating

functions

10(s) = M ln<exp(slr|j2)H> 0s > 0 (6. D.4)

Page 152: ATMOSPHERIC PROPAGATION

-152-

'Pl(s) = M ln<exp(slr12) IH1 >s < 0 (6.D.5)

where s is a real variable, approximate expressions for the false

alarm probability PF and miss probability PM = 1 - PD are

PF = x oi(s ) - s (s ) + s20 (S )/2] Q[/vlgso) so]F 0 0 0 0 0 00 0

P = exp[il(sl) - sl S + l (Sl )S 21 QE-vil(s, ) sl

(6.D.6)

(6.D.7)

for M 5. In (6.D.6), (6.D.7) dot denotes differentiation with

respect to s, Q(-) is the complemented error function

.co

Q(y) = exp(-x2 /2)

/- r

dx (6.D.8)

and sl' s1 are solutions to the equations

My = O(s)

my = l (S)

(6.D.9)

(6.D.10)

respectively. The approximation used to derive (6.D.6), (6.D.7) was

an application of the central limit theorem [15]. Specifically,

Page 153: ATMOSPHERIC PROPAGATION

-153-

Mthe sum j tn 2 is taken to be a Gaussian random variable even

i=lthough the Ir.! i = 1,2,...,M are not Gaussian. Hence, the requirement

M 5.

For the PF calculation H is true so that r2= tn.!22 , a

unit mean exponential random variable, hence

y1(s) = -M ln(l - s) s < 1 (6.D.ll)

Equation (6.D.9) is easily solved to give the threshold

1 (6.D.12)1s 0

with approximate false alarm probability

M M(3s /2 - s sP (1 - s )~ exp Q 0 (6.D.13)

F o (L - s0) L ~"

Equation (6.D.13) can be inverted numerically to give y as a function

of M and PF* In Figure 6.6, we show M vs. y for PF = 1012. Clearly

a modest amount of pulse integration leads to a drastic reduction in

threshold.

The function v1l(s), and hence the detection probability PP'

depends on the target/atmospheric scenario. We present our results

as a series of examples.

Page 154: ATMOSPHERIC PROPAGATION

-154-

100

90-

80-

70-

10 -12PF

60-

M 50-

40-

30-

20-

10-

0 5 10 15 20 25 30 35Y

116599-N

Figure 6.6: Threshold y vs. number of pulses necessary to

maintain PF 0-12

Page 155: ATMOSPHERIC PROPAGATION

-155-

Case 1. jxj 2 Nonrandom

If 1x2 = CNR is known the moment generating function is

given by [31]

y (s) = M CNR T - ln(-s) s < 1 (6.D.14)

This result is useful in the jxf 2 random cases as we can average

exp(vil(s)/M) with respect to the statistics of Ix12 to find p1 (s).

Case 2. Single-Glint Target in Turbulence

For a single glint target T(p') = e a T (P') satisfying

(4.D.1) - (4.D.3) we find that [31]

pl (s) = M ln s Fr{- s CNR, 0; 2r s < 0 (6.D.15)

where Fr(a,0;c) is the lognormal density frustration function [1,33,34].

Saddle-point integration techniques for Fr(a,0;c) that are in the

literature [27,33] permit rapid accurate numerical evaluation of -pl(s)

and, in turn, of PD.

Case 3. |x12 Exponential

Here we consider the case of Jxj 2 = CNR v where v is a unit

mean exponential random variable. As before, this characterization

applies to the three scenarios:

Page 156: ATMOSPHERIC PROPAGATION

-156-

.1. Free space, speckle target.

2. Low visibility, large speckle target.

3. Low visibility, resolved glint target.

For this case, we have [31]

Pi(s) = -M ln[l - s(l + CNR)]

Case 4. Resolved

For a target with

becomes [31]

Speckle Target in Turbulence

radar return described by (4.D.9) yi(s)

ii(s) = M ln { dU p(U)[1 - (1+ CNRsr e2U)s]l} s < 0 (6.0.17)

with

(6.D.18)Pu (u) = exp[-(U + a 2 )2 /2 C 2]

/2 ar2

Case 5. Ix2 A Product of Two Exponential Random Variables

The case of a small glint target in low visibility is considered

here, jx2 = CNR vw, where v, w are IID unit mean exponential random

variables. We have that

(S) = M CNR - ln (-sCNR) + lnSs - 1sCNR

ey- dt s < 0 (6.D.19)

s < CNR + 1(6.D.16)

Page 157: ATMOSPHERIC PROPAGATION

-157-

As in the single pulse M = 1 case we are most interested in

case 5 above and in contrasting it with case 2 as they both deal with

small, single glint targets. Numerical work on case 2 has appeared

in [35] and will be used for comparison.

In Figure 6.7, PD is plotted vs. CNR for Ix12 a product of

two exponentials, PF = 10~4, and several values of M. In Figure 6.8,

is a similar set of curves for PF = 12. These two figures indicate

the performance improvement obtained through pulse integration.

In Figure 6.9, the number of pulses M necessary to achieve a

performance of PF = 10-12 ,D = 0.99 is plotted vs. CNR for a glint

target in free space, in turbulence for two values of a2 , and in lowX

visibility. This figure indicates that, for equal CNR a glint target

seen through scattering conditions can be detected more easily than

when seen through saturated scintillation a2 = 0.5. Also atmosphericX

pulse-integration performance for large M is very near that for free

space as only 1 or 2 more dB of CNR is necessary to achieve

PF = 10-12 PD = 0.99 when M 20 in the atmosphere except in heavy

turbulence.

In Figure 6.10, we again plot M vs. CNR, for a glint target

in bad weather, PD = 0.99, and two values of false alarm probability

Page 158: ATMOSPHERIC PROPAGATION

-15-

I I

M =30

-/- M =10-

|X|2 2 VW

- ~ P 1=M1 4

F

10 20 30

CNR (dB)

Figure 6.7: Single-pulse and multipulse detection probability 2-

vs. CNR for a glint target in bad weather. PF= 104

throughout.

99.99

99.9

99 F-

95

90

70

50

30

10

5

CL

C.

4).

01

0.1

U 40

I II I

Page 159: ATMOSPHERIC PROPAGATION

-159-

M = 30

jVW12

S>V Iw

100

CNR (dB)

Figure 6.8: Single-pulse and multipulse detection probability

vs

PF

CNR for a glint target in bad weather.

1 12 throughout.

99.99

99.9

I

99 K-

95

90

70

4,'a 50

30- PI <

10

10

5

0.1

0.01

M 1

20-20 -10

v

I I

Page 160: ATMOSPHERIC PROPAGATION

-160-

100

90P = 0.01 P= 0.99

M D

PF = 10-12

70 -

bu GLINT TARGETLOW VISIBILITY|>12 2> VW

M 50

40-

30-

GLINT TARGET 2= 0.0520 -X

GLINT TARGET, FREE SPACE

GLINT TARGET .2 = 0.510- x

0 10 20 30 40 50 60 70CNR(dB) 116596-N

Figure 6.9: The number of pulses M necessary to achieve PF = -12

pD = .99 vs. CNR for a glint target in free-space, two

turbulent fluctuation levels, and bad weather.

Page 161: ATMOSPHERIC PROPAGATION

-161-

M

F

--- p =

MP F

0.0110-7

GLINT TARGET,

1 lxi2 <x

.

P =0.99D

0.01 P = 0.9910-12

LOW VISIBILITY2

> vw

10 20 30 40 50 60 70CNR (dB)

116595-NFigure 6.10: The number of pulses M necessary to

achieve PD = .99 and two different false

alarm probabilities vs.

in low visbility.

CNR for a glint target

rv-~I~i I

90

I1*1I

II

60K

M 50

40

30

20

10

0

so -

70k-

Page 162: ATMOSPHERIC PROPAGATION

-162-

PF -10 12. From this figure it is clear that for just 1 or 2

extra dB of CNR a 5 order of magnitude improvement in false alarm

probability can be achieved without sacrificing detection sensitivity.

In Figures 6.11, 6.12, PD is plotted vs. CNR for M = 10, 15

respectively, PF = 10-12, a glint target in several turbulent

atmospheres, and the same target in a turbid atmosphere. Again we

see that for equal CNR the target in the turbid atmosphere is easier

to detect than for saturated scintillation a' = 0.5.X

The implications of the above results cannot be properly

accessed without considering resolution and CNR in bad weather. From

(4.B.11), (4.E.4), etc., we have that in bad weather, both resolution

and CNR are degraded from their corresponding free-space values while [5]

in turbulence these same quantities remain essentially unchanged as

compared to free-space. What this means is that targets which would

be resolved in turbulence (and free-space) may not be in bad weather.

Further, the CNR available to the radar for the same target, viewed

through the two limiting cases of clear and scattering atmospheres,

will be much smaller in the latter case as compared to the former.

Hence, the difficulty in detecting a target in bad weather is expected

to increase as compared to in turbulence. In the next chapter, CNR

and resolution curves are presented for typical systems and various

targets in bad weather. These curves, coupled with the results just

presented, can be used to access target detectability in inclement

weather.

Page 163: ATMOSPHERIC PROPAGATION

-163-

S=0.01-x.-2 A

-

-x

- I

1XI )V W -I

30 -

PF = 10- 12

M =10

10 20

CNR (dB)

Figure 6.11: Ten pulse detection probability vs.

target in several turbulent atmospheres and a

glintiscattering

atmosphere. PF = 10-12 throughout.

99.99

99.9

99

=0.5

95

90

70

50

Ca,C.)a)0.

0a-

10

5

1

0.1

0.010 30 40

CNR for a

I I1i I

Page 164: ATMOSPHERIC PROPAGATION

99.99

99.9

99

2'2- X .1

- ,/

-- I

I2II >vw

2x = 0.5X

30-

10 KF

-1012

M = 15

01 I-

Ui 1u

CNR (dB)

Figure 6.12: Fifteen pulse detection probability vs.

a glint target in several turbulent atmospheres

and a scattering atmosphere. PF = 0-12 throughout.

95

90

70

50

C

U

0.

00~

5 F

1

0.0120 30 40

CNR for

I I I

01

L I

Page 165: ATMOSPHERIC PROPAGATION

-165-

CHAPTER VII

SYSTEM EXAMPLES

The preceding chapters have pointed out three primary

contributions to the optical radar return:

(1) the carrier-to-noise ratio CNR due to MFS propagation,

which as discussed in Chapter II and Appendix A, disregards

the unscattered portion of the light beam;

(2) the extinguished free-space carrier-to-noise ratio

CNR0 e-2tL, which is due to the unscattered light; and

(3) the normalized backscatter power <1x%(t)1 2 > from scatterers

in the propagation path between radar and target which, in

the context of an imaging or detection application is

undesirable.

The main theoretical development of this thesis has been the scattered-

light MFS propagation theory as applied to an optical radar. In this

chapter, we introduce two representative hypothetical radar systems

and examine, for a variety of target/atmospheric scenarios, the

relative strengths of the above mentioned radar-return contributions.

Most significant here are examples which indicate where the MFS return

is dominant, for these establish regions of applicability for our

theoretical work.

Page 166: ATMOSPHERIC PROPAGATION

-166-

The two hypothetical radar systems we employ in this chapter

are a CO2 laser radar, and a Nd:YAG laser radar. Their essential

parameters are summarized in Tables 7.1 and 7.2, respectively. The CO2

laser radar parameters are chosen to closely match a typical existing

system of that type, such as the mobile, Lincoln Laboratory infrared

radar mentioned previously. The Nd:YAG laser radar parameters are

chosen purposely to give an example of a system which can exploit

scattered light. Indeed, the system as described here is on the edge

of current technology and hence, at this time, may be unrealizable. It

provides, however, a context in which MFS propagation is often the

dominant target return mechanism.

The CNR results of Chapter IV, Section E all assume a uniform

scattering profile between radar and target. The examples in this

chapter will indicate that neither the C02 system nor the Nd:YAG

system will be able to make use of scattered light in this situation

because of the dominance of the extinguished free-space power. It

turns out, however, that there are several interesting situations in

which the MFS power is dominant, at least for the Nd:YAG system. These

situations arise when the propagation path does not possess a uniform

scattering profile between radar and target, but rather the scattering

is concentrated within a layer comrpising a small to modest fraction

of the total path length. Such situations. occur in the context of an

airborne radar searching the ground through a cloud or fog cover, or a

ground-based radar searching the sky, again through a cloud or fog

Page 167: ATMOSPHERIC PROPAGATION

WAVELENGTH X

PHOTON ENERGY hv0

PEAK POWER PT

PULSE DURATION tp

BEAM DIAMETER 2PT

DETECTOR QUANTUM EFFICIENCY r

1.87 x 10-20

10 kW

100 ns

13 cm

0.9

46

Table 7.1: CO2 Laser Radar System Parameters

-167-

10.6 -pm

Page 168: ATMOSPHERIC PROPAGATION

WAVELENGTH A

PHOTON ENERGY hv 0

PEAK POWER PT

PULSE DURATION t

BEAM DIAMETER 2PT

DETECTOR QUANTUM EFFICIENCY rn

1.87 x 10 9 1

10 kW

100 ns

0.5 cm

0.9

Table 7.2: Nd:YAG Laser Radar System Parameters

-168-

1.06 pm

Page 169: ATMOSPHERIC PROPAGATION

-169-

cover.

Two layered profiles are considered in this chapter. The first

consists of a layer of scatterers of thickness Ls sandwiched between

two free-space layers of thickness Li and Lo, where the distance from

the radar to scattering layer is L the distance from the scattering

layer to the target is LO and the total target range is L = Li+ Ls+ L .

This situation is depicted in Figure 7.1. The second layer geometry

consists of scatterers of thickness. L placed near the target with L.

meters of free-space between the radar and the scatterers, so that the

total target range is L = Li + Ls. This second situation is depicted

in Figure 7.2. The formulas of Chapter IV, Section E do not apply

without modification to the layered scattering scenarios. Fortunately,

necessary modifications are not too extensive. Hence, we will indicate

what the appropriate modifications are as the examples are discussed.

Finally, before delving into our calculations we must indicate

values for the atmospheric parameters. The numbers we use are

summarized in Tables 7.3 and 7.4, and are typical [18,37,40]. The

numbers cited for haze are used in all examples in which a uniform

scattering profile is assumed. These examples are intended to predict

how the radar systems would perform in a terrestrial situation. The

numbers cited for a cloud are used in all examples in which a layered

scattering profile is assumed. These examples are intended to predict

radar performance in air to ground or ground to air scenarios.

Page 170: ATMOSPHERIC PROPAGATION

FREE-SPACE

L

SCATTERINGLAYER

L

L

FREE-SPACE

Figure 7.1: Geometry for a single scattering layer between radar and target.

RADAR TARGET

Is

Page 171: ATMOSPHERIC PROPAGATION

FREE-SPACE

L.

L

SCATTERING TARGETLAYER

Ls

Figure 7.2: Geometry for a scattering layer near the target.

RADAR

-4

Page 172: ATMOSPHERIC PROPAGATION

-172-

Extinction Coefficient St (km1 )

Modified Scattering Coefficient ' (km~ )

Backscattering Cross Section Per Unit

Volume (km1 )

RMS Forward Scattering Angle 6F (mrad)

Atmospheric Parameters at CO2 Laser Wavelength

Haze

.005

.0005

.0005

50

Cloud

10

4

.2

50 -

Table 7.3:

Page 173: ATMOSPHERIC PROPAGATION

-173-

Extinction Coefficient t (km~1 )

Modified Scattering Coefficient

s' (kmi )

Backscattering Cross Section

Per Unit Volume (km 1 )

RMS Forward Scattering Angle

eF (mrad)

Table 7.4: Atmospheric Parameters at Nd:YAG Laser Wavelength

Haze

.07

.05

.01

10

Cloud

20

17

5

1

Page 174: ATMOSPHERIC PROPAGATION

-174-

A. Examples

Figure 7.3 gives plots of the normalized backscatter power from

a haze vs. delay t, Equation (4.H.12), for both the CO2 and Nd:YAG

systems. Since the integration interval in (4.H.12) is ct p/2 = 15 m,

we can regard <1 xb(t)1 2> as the backscatter return from range L = ct/2

so that, for example, t = 10 psec corresponds to L = 1.5 km and

t = 70 ypsec corresponds to 10.5 km. Hence, Figure 7.3 covers the entire

target range of 1 km < L < 10 km considered. What we see from this

figure is that in a haze <!xg(t)(2 > is always less than 20 dB for the

CO2 system, and below 0 dB for the Nd:YAG system beyond t = 10 ypsec

(L = 1.5 km). We will see later that the target return power for a

resolved speckle target will dominate this backscattered power for both

systems, so that < x-(t)1 2 > can be disregarded for uniform haze

profiles.

Before presenting the above mentioned target return result,

we shall give our second and last backscatter example. Following a

procedure similar to that of Chapter IV, Section H, we can use single

scatter theory [28] to derive the following expression for the maximum

normalized backscatter power from a scattering layer L meters from the1

radar

< 1Sbi) 2> 'nI t 0P T 2 cobP<h, x-(L P22 h2 dz (7. A.1)

o t T -L. c- t 1 + 2 z22 2p 27T 2

Page 175: ATMOSPHERIC PROPAGATION

70

50

30

- V-__

-10.

t -7

CO 2 System

Nd:YAG System

~1

LL-~--

20 40 60

t(iisec)

Figure 7.3: Normalized backscatter power from a uniform scattering

profile vs. t.

-- fi1jf~z~f~~

A

C%4

V

10

-300

-----------

-

Page 176: ATMOSPHERIC PROPAGATION

-176-

This maximum normalized backscatter return results when the pulse of

length ct p/2 (meters) has just entered the scattering layer. In this

case, extinction can be neglected as exp(-2 t ct p/2) 1 1 for the pulse

duration considered. Plots of the maximum backscatter are shown in

Figure 7.4 for both radar systems, where the atmospheric parameters

assumed are those of a cloud. As one might expect this backscatter

return is significantly larger than for a haze. However, if the radar

uses a range gate that effectively closes the receiver aperture until

the desired moment, this backscatter power can be combatted. As the

laser pulse propagates through the cloud the backscatter will decrease

because of extinction. A good approximation to the backscatter power

from a distance Ls into the scattering layer is (7.A.1) multiplied by

exp(-2 tLs). If this quantity is smaller than either CNR or

CNR 0 exp(-2t L s) then the backscattered power can be excluded via range

gating. Later calculations will show this to be the case.

The case of a resolved speckle target seen through a haze

(uniform scattering profile) is summarized in Figures 7.5 and 7.6. Figure

7.5 gives the CO2 system extinguished free-space and MFS CNR, where the

target mean square reflection coefficient r= 0.5. Figure 7.6 gives the

same quantities for the Nd:YAG system with the same value for r. In both

cases the extinguished free-space return clearly dominates the MFS return

so that the target return is best described by (4.D.9) and resolution is

given by the free-space result. Also by comparing the extinguished free-

space curves with the backscatter curves of Figure 7.3 we see that

Page 177: ATMOSPHERIC PROPAGATION

-177-

70

50

30

-77-

~Iit -~

10

-~ ~ -iL::yiiI$:T7!1i7:7f:::.: :7: ::Z7 :-

-~:c:::r:V

0 2

CO2 System

Nd:YAG System

4

Li (km)

Figure 7.4: Maximum normalized backscatter power from a scattering layer

L. meters from the radar.

~K ~

^M

A

C\J

><P 10v

6

:f

-30

Page 178: ATMOSPHERIC PROPAGATION

-178-

100

80

60

40

20

0

L N

CNR

____- -- ~---- -- T

0 4 8 12

L(km)

Figure 7.5: Extinguished free-space and MFS resolved speckle target CNR

vs. target range for the C02 system and a uniform scattering

profile.

L 4

L-:rr_ _=

'77

0 - 2 tLCNR et

Page 179: ATMOSPHERIC PROPAGATION

0

-179-,

80

60

40

-o

20

0

-20

-= o -2 tLCi CN R et

~I-v I ~ -~ -~

§1272:.

CNR

4 8 12

L ( km)

Figure 7.6: Extinguished free-space and MFS resolved speckle target CNR

vs. target range for the Nd:YAG system and a uniform

scattering profile.

LA L

77_

7 7 7:: L=

7 1 ----- -

L'=7 z__

Page 180: ATMOSPHERIC PROPAGATION

-180-

backscatter is insignificant in a haze for a resolved speckle target.

The dominance of the extinguished free-space return can be understood

if one considers that for either radar system, even at target range

L = 10 km, the beam has not propagated through a single optical thickness.

That is, StL < 1.

For a resolved speckle target with a thin layer of scatterers

between it and the radar we might expect, because the optical thickness

is much larger than in the previous example and many more scatterers

are encountered, that the MFS return would dominate. For the Nd:YAG

system, since the single scatter albedo is nearly unity at this wavelength,

we shall see that this is in fact the case. If the speckle target is

as in Figure 7.1 with L. >> Ls, L0 >> Ls, it is a simple matter to show

that the appropriate results of Chapter IV (i.e., (4.B.7)-(4.B.9),

(4.E.7), (4.E.8)) all hold with p0 given by

P = 2 2 0 3 kol(7.A.2)o ' Ls O k2

in place of (2.C.3). With this in mind, the case of a resolved speckle

target of mean square reflection strength. r = 0.5 and a thin scattering

layer between radar and target is summarized in Figures 7.7-7.9.

Figure 7.8 shows the CO2 system extinguished free-space and MFS CNR

for this scenario with L. = 2 km, L0 = 1 km vs. scattering layer

thickness Ls. These same quantities are shown for the same conditions

Page 181: ATMOSPHERIC PROPAGATION

100

80

60

~Th~-- 4k V7t~z~

CO System2

Nd:YAG System~z~zc~z~Z2V77

0 200 400

SLffi~~F I

600

L s (m)

Figure 7.7: Atmospheric beamwidth for a single layer scattering

profile vs. layer thickness.

P~7iZEZ2Z27

E

S.-

40

20

11

Page 182: ATMOSPHERIC PROPAGATION

-182-

80

60

40

tR\7

t-Z

0

CNR 0 e-2tLs

20--

CNR

-20 ' --C 200 400 600

L s (m)

Figure 7.8: Extinguished free-space and MFS resolved speckle target

CNR for the CO2 system and a single scattering layer

vs. layer thickness.

- _7

--7::_

Page 183: ATMOSPHERIC PROPAGATION

200 400 600

Ls (m)

Figure 7.9: Extinguished free-space and MFS resolved speckle

target CNR for the Nd:YAG system and a single

scattering layer vs. layer thickness.

70

50

30

10

-10

-30C

- i __

- * CNR

CNR0 - t s

- 7 --- --- --- ----

Page 184: ATMOSPHERIC PROPAGATION

-184-

and the Nd:YAG system in Figure 7.9. For the CO2 system we see

that the extinguished free-space beam dominates. But for a

scattering layer thicker than 150 m the MFS power dominates so

that it is appropriate to describe the target return by (4.E.29).

The backscatter can be disregarded as (considering the geometry

of Figure 7.1) it can be easily range gated out.

Besides power return we must also consider resolution, i.e.

whether or not the beam lies entirely on the target, as assumed

for this example. Figure 7.7 shows the atmospheric (MFS) beamwidth

for both radar systems, and L = 2 km, L0 = 1 km vs. scattering

layer thickness Ls. Since the CO2 radar return is given by the

extinguished free-space result, beam size in this case is given by

the free-space result and the upper curve can be disregarded. But

since for the Nd:YAG system the MFS return dominates for Ls > 150 m

the beamwidth we must consider is the atmospheric limited result

shown. Note that if the target is 10 m or larger in diameter it

is resolved and the curves, Figure 7.9, apply.

For a resolved speckle target with a scattering layer near

the target (Figure 7.2) we expect that forward-path beam spread

loss and receiver coherence loss should be less severe than in the

previous examples thus increasing the MFS return relative to the

extinguished free-space return. For this case it can be easily

shown that Equations (4.B.7)-(4.B.9), (4.E.7), (4.E.8) are again

appropriate if in place of (2.C.3) we use

Page 185: ATMOSPHERIC PROPAGATION

-185-

p 2 ' (7.A.3)0 'Ls e2k 2 Lsl

An example of the use of the Nd:YAG system in this situation is

summarized in Figures 7.10 and 7.11 where we have assumed L. = 1 km.

Figure 7.11 shows both the extinguished free-space and MFS CNR for

this situation. As the MFS return dominates we need to consider the

atmospheric beamwidth shown in Figure 7.10. For a scattering layer

thickness of less than 400 m a target of 1 m radius or more is

resolved so that the target return is described by (4.E.29). The

location of the scattering layer, in this case, is such that the

backscatter power must be carefully considered. From Figure 7.4,

the maximum normalized backscattered power is 31 dB. In Figure 7.11,

we plot the product of this maximum backscattered power and the

extinction factor exp(.-2t Ls ) vs. Ls. It is clear that if we range

gate properly we can receive the true MFS target return. For example,

if our target is 300 meters inside the scattering layer then

CNR = 23 dB. From Figure 7.11, the backscatter return falls below

this level at Ls = 50 m so that if we open the range gate

2 x 250 m/c = 1.67 iisec or less before the target return arrives we

will receive the target return and not the backscatter power. This

should not be too difficult if we have apriori knowledge of where the

cloud begins.

Page 186: ATMOSPHERIC PROPAGATION

0

-186-

1.0

0.8

0.6

-zS.- 0.4

0.2

0 200 400 600

Ls (m)

Figure 7.10: Atmospheric beamwidth for a scattering layer near the

target and the Nd:YAG system vs. layer thickness.

------------

-- --------

Page 187: ATMOSPHERIC PROPAGATION

-187-

70

50

30

T0

-10

-CNR

7 7_

CN RO e

2: - tF7e

-30C 200 400 600

L s (m)

Figure 7.11: Extinguished free-space and MFS resolved speckle target CNR

for the Nd:YAG system and a scattering layer near the target

vs. layer thickness.

Page 188: ATMOSPHERIC PROPAGATION

-188-

For an unresolved glint target in the situation depicted

in Figure 7.2, CNR equations similar to those for the same target

in a uniform scattering profile apply. These are

CNR = CNR 0

1/3(XL/Trpin )21 +

(XL/2rpT 2P +T

2

er2

+ c

rlP tpCNR 0 = T p

h\) 0

4r 4s

[(XL/27r T2 + pZ]

as before and

rb = ( XLw 2+(L/21rpT)2 +prb ( L/7Tpin) 2 + +A2rP

(AL/7T) 2 rb

r [ (AL/Tpout )2 + p2] - [p - (ALrpmid 2

where

(7.A.3)

(7.A.4)

r2c

(7.A.5)

(7.A.6)

Page 189: ATMOSPHERIC PROPAGATION

-189-

p 2 2 L 2 (7.A.7)out ' k2 e2 L3 - Ls F1

p 2 L (7.A.8)in k2 e2 L3

sF s

P2 2 L2 (7.A.9)mid 6' k2 82 3LL - 2L3s F s s

An example of the use of the Nd:YAG system in this situation is

summarized in Figures 7.12, 7.13 for L. = 1 km. Figure 7.13

shows extinguished free-space and MFS CNR for a glint target of

radius rs = 0.25 mm and mean square reflection coefficient r = .5.

For scattering layer thickness L > 300 m the MFS power dominates.

But at this point, CR0 e s = CNR = 6 dB. Hence the MFS power is

difficult to "see" for this target. The maximum backscatter power

is again 31 dB and just as in the last example can be range gated

out.

The small-glint detection theory analysis in Chapter VI

(i.e. the Jxj 2 = < 2> vw case) requires that the MFS return be

dominant. It also requires that the target radius rs be less than

the field coherence length rc. In Figure 7.12, rc is plotted for

this case vs. scattering layer thickness. We see that for Ls > 300 m,

where the MFS return dominates the condition rs S rc holds with

approximate equality.

Page 190: ATMOSPHERIC PROPAGATION

-190-.

i7I 7

---- ----

I 4~

7. -7--.7- -- ------

0 200 400 600

L S(in)

Figure 7.12: Field coherence length for a scattering layer near the

target and the Nd:YAG system vs. layer thickness.

1.0

0 .Q1008

0.0 006

E

C-)5-

0.0 004

0.0002

0.0

z

77

Page 191: ATMOSPHERIC PROPAGATION

-191-

-- - - - T- -L r

T__-- - - - -

t. -' - -~ --

H7 1-:t 7

,7, 7 ,7- + 7_

=tE=....... ......

::i 7

--- CIN R

1171 -~-x\ - X j1Z2Ft1"

.-m -- M K -26tLs.. ....CINR e

400 600

<K.~jL2V e2BtLs30 x 2e.b

- 0 200

L s(M)

Figure 7.13: Extinguished free-space and MFS unresolved glint

target CNR for the Nd:YAG system and a scattering

layer near the target vs. layer thickness.

70

50

Ju

7-

11~1 I-- j~~J

Page 192: ATMOSPHERIC PROPAGATION

-192-

The above examples indicate the regions of applicability

of the various target return models considered in this thesis in

terms of two laser radar systems. We observed that for resolved

speckle targets and the Nd:YAG system, the MFS power can easily

domainte all other radar returns. For the CO2 system, we found that

it is appropriate to treat low visibility weather as a purely

extinction phenomenon. Hence, resolution of the Nd:YAG system can

be degraded in bad weather, from the free-space result while this

degradation does not exist for the CO2 system. The reason for this

is simply the higher single scatter albedo at the shorter wavelength.

That is, nearly all light incident on a scattering particle at the

Nd:YAG wavelength is scattered while at the CO2 wavelength about

half the power is absorbed.

It is clear that in an imaging application any radar system

that makes use of scattered light is going to have degraded

resolution as compared to free-space. This leads one to the

conclusion that scattered light is most useful in a detection

application. At this point, though, it is not clear whether one

would want to design a system specifically to use scattered light.

That is, the question of whether one detects targets more easily

with a system designed, as in the Nd:YAG system above, to have

the MFS return dominate or, as in the CO2 system, to have the

extinguished free-space return dominate has not been answered. In

order to answer this question the implications of the CNR results,

Page 193: ATMOSPHERIC PROPAGATION

-193-

as in the examples of this chapter, on the false alarm/detection

probabilities needs to be investigated and the tradeoffs evaluated.

Page 194: ATMOSPHERIC PROPAGATION

-194-

CHAPTER VIII

SUMMARY

In this thesis, we have examined the use of a heterodyne-

reception optical radar in both imaging and target detection

applications. As noted early in this work, such systems may be

severly limited by the stochastic nature of atmospheric optical

propagation; that is, by turbulence, absorption, and scattering. We

began the thesis by presenting a mathematical system model which

incorporates not only the statistical effects of propagation through

either turbulent or turbid atmospheric conditions, but also target

speckle and glint and local oscillator shot noise. We later

augmented this model to account for beam wander induced fluctuations,

whether due to radar aimpoint jitter or turbulent atmosphere beam

steering.

Once this model was established we used it first to analyze

the radar in a scanning-imaging application, Most important here

were the issues of resolution and signal-to-noise ratio. In

previous work turbulent atmosphere resolution had been shown to be

the same as free-space. We showed, as was expected, that turbid

atmosphere resolution is degraded from the free-space value. We

found, due to the statistical nature of the local oscillator shot

noise, that the signal-to-noise ratio depended on two quantities:

Page 195: ATMOSPHERIC PROPAGATION

-195-

SNRSAT and CNR. The former being the SNR limit set by target return

fluctuations, the latter being the average target return to shot

noise power ratio. In turbulence, these quantities had previously

been evaluated for a number of interesting targets. We proceeded to

evaluate these same quantities for inclement weather. In the course

of doing so, complete statistical characterizations of the radar

return were developed for several bad weather situations, as had

been done previously in turbulence. In both turbulence and bad

weather, we endeavored to interpret CNR and SNRSAT results in terms

of intuitively pleasing descriptions of target interaction and

atmospheric propagation. These interpretations greatly enhance our

understanding of the mechanisms that degrade optical radar

performance.

The second major use we made of our system model was in

analyzing the detection capability of the radar. For each complete

statistical description of the radar return developed earlier, a

separate detection performance analysis was required. We concentrated

on the case of a small glint target in bad weather, as this

represented new work.

We were also concerned with verifying the theoretical

statistical characterizations of the radar return. Due to the nature

of the available experimental setup we were limited to verifying our

beam wander and turbulent atmosphere models. We found very good

correspondence between theory and data in many cases. Of particular

Page 196: ATMOSPHERIC PROPAGATION

-196-

interest here was the verification of the lognormal character of

the atmospherically induced fluctuations.

Our last task involved the establishment of regimes of

validity for the various target return models. This was done in

terms of two hypothetical optical radar systems and involved comparing

relative values of three quantities: The extinguished free-space

radar return; the MFS radar return; and the atmospheric backscatter

return. It was shown that some regime existed wherein every target

return model was valid.

In the future, several topics which extend and adjoin this

work might be investigated. These include: continuation of theory

verification work; a similar type of analysis, as found herein,

performed on a direct detection system which might make better use

of scattered light; generalization of this work to include doppler

shift-moving target indication; use of an array of detectors in

place of the single detector assumed in this work to make better

use of scattered light; and an investigation with other values for

system parameters than in this thesis (for example, unequal transmitter

and receiver aperture diameters). All of these topics are important

for a more thorough understanding of optical radar potential, and

should all be investigated

Page 197: ATMOSPHERIC PROPAGATION

-197-

APPENDIX A

DERIVATION OF THE MUTUAL COHERENCE FUNCTION (MCF)

In this appendix, the details of the derivation of the

multiple forward scattering (MFS) propagation model, Eq. (2.C.2) are

given. We begin by defining the specific intensity I(r,f) and give its

governing equation, the equation of transfer, which is essentially a

statement of energy conservation. This equation is specialized to the

case of a collimated beam and the specific intensity is then shown to

be related to the MCF by a Fourier transformation. The specialized

equation of transfer is then Fourier transformed, and its solution,

the MCF, given. We then apply this result to real scattering

atmospheres and derive the multiple forward scatter (MFS) propagation

theory.

Consider a flow of wave energy at some point r in a random

medium. For a given direction defined by the unit vector U we can find

the average power flux density per unit solid angle. This quantity

I(F,!) is called the specific intensity [28] and has units W m-2 sr~ .

Operationally, we can say that the amount of power dP flowing within a

solid angle d0 and through an elementary area da oriented in the

direction of the unit vector ? is (Figure A.1)

Page 198: ATMOSPHERIC PROPAGATION

/ ...

e / d/

Figure A.: Geometry relating to the definition of specific intensity.

0

00

1

J01

I I b

da

Page 199: ATMOSPHERIC PROPAGATION

-199-

dP = I(r,2) cos(6)da do (A.1)

The equation that governs the evolution of the specific intensity is

the equation of transfer [28] and for a sourceless medium is

Q*V I ,2 t I~r~ 4wf5 dZ' p( T,Q) I(r,Q') = 0 (A.2)

In (A.2), t(s) is the extinction (scattering) coefficient (m~1 ) and

p(f,f') is the single particle phase function normalized to satisfy

dff p(2,Q') = 1 (A.3)47r

Equation (A.2) can be interpreted as saying that the change in power

flowing in the 0 direction due to propagation in a random scattering

medium is equal to the power scattered into that direction from all other

directions less the power absorbed and, hence, is a statement of

energy conservation.

If we now assume that all light of interest is propagating

nominally in the +z direction, so that we can say

S= (, /l - [Ij2) ~ (il) (A.4)

where

Page 200: ATMOSPHERIC PROPAGATION

-200-

S = (s ,s ) (A.5)

then (A.2) becomes

s v- I(P, ,z) + Z I(Fiz) + pt(,isz) -- TS ds' p(S-') I(P,' ,z) = 0

(A.6)

In (A.6), r = (P,z) where P = (x,y) is the coordinate vector transverse

to the direction of propagation, Vp = x 3/ x + y 3/3y is the two

dimensional del operator and the phase function is assumed to be a

function only of the difference in the output direction and the direction

of the incident wave. Also p(s) is assumed to be sufficiently narrow

that the limits on the integral in (A.6) can be extended to infinity.

To relate the specific intensity to the mutual coherence

function consider Figure A.2. The power incident on the detector is

easily shown to be

r ~r p P 2,!L) k -P = jd d 2 Jdp' exp(-j y p'-(p -p 2 ))circ(2 1- -- p /dR

-circ(2152 o/d) cir(21p' - pDI/d) (A.7)

where r( P, 2,z) = <u(Pi,z) u*(P2,z)> is the MCF, the p, and P2

integrals are over the (z = L) receiver plane, the p' integral is over

Page 201: ATMOSPHERIC PROPAGATION

RANDOM,TIME INDEPENDENT SPATIALLY VARYINGPROPAGATION MEDIUM

u~,L) ....

Z=O

I- DIAMETER dRFOCAL LENGTH fCENTERED AT

Z=L

0

DETECTORAREA A = ird 4CENTERED AT =f i

D

§0

Z= L+f

Figure A.2: Geometry for relating the specific intensity and the mutual

coherence function.

C3I

Page 202: ATMOSPHERIC PROPAGATION

-202-

the (z = L + f) detector plane and circ(2[pl/d) is a circular pupil

function defined by

circ(2I5I/d) =

I

0

1PI < d/2

(A.8)

elsewhere

If we assume the detector and AD is equal to a diffraction limited

field of view

AD = (Xf/dR) 2 (A.9)

the detector plane integral can be approximated so that (A.7) becomes

=I A C)P dp C dp Pd , ) C VL) exp(--jk 's Pd )cic_1cT -o R

2- -circ(2I-pc + 2 -P 0 /d R) (A.10)

where r' is the MCF in terms of the sum and difference receiver plane

coordinates pc = + 2)/2 and Pd = l - P2. As the field u(P,z)

has propagated through L meters of the random, time independent

spatially-varying propagation medium it is reasonable to assume that

'(c' d,L) = -'(p ,pdL) over the receiver aperture. Further assuming

that r'(. ,Pd,L) is narrower than the receiver aperture in Pd' i.e.,

Page 203: ATMOSPHERIC PROPAGATION

-203-

'(PoPd, L) Z 0 (A.11)pd > Pa

where pa << dR/ 2, equation (A.10) becomes

A 7rd2 -P =R d d F o'd,L) exp(-jk s-pd) (A.12)

If we write this same detected power in terms of the specific intensity

we have

rd2 AP = I(P ,s ,L) R D

o 0 4 2(A.13)

Combining (A.12), (A.13), we have that r' and I are related by a

Fourier transformation

I(f oL) =i drdfr dL) exp(-jk - (A.14)

T' (PC') d, L) = s d I(-Pc' 5s0,L) exp(+jk so -Pd) (A.15)

If we Fourier transform (A.6), we then have the differential

equation that the MCF must satisfy

Page 204: ATMOSPHERIC PROPAGATION

-204-

1 d v I,z +) + 5 -'(ic'Pd,z) + ,P ,z)Jk '7 d -7cr(c~dz 9 cPdt cpd~

- s(Td) F'Cpc' d,z) = 0

where P(Pd) is the Fourier transform of the phase function p(s)

jks *dP(Pd jds 0p(so) e 0 d

Equation (A.16) has been solved [18,28] and

'(pc' dZ) = j dpd r '(', ,0)e jdp Jc' ' (Xz) 2 e p Pd C c

rr-z rz + s d- exp - S) dS - I s(s) 1 - P zpl-]+pd ds1

(O a.o

(A. 18)

which implies

= ho(p ,5 ) ho(, 2 '2I)

{ L ( L

-exp - a(z) dz - BS (Z) 1 - P [ 1 - + zp dz

.0 .0 (A. 19)

(A.16)

(A.17)

<hL (,p -) h*(-2,p )>

Page 205: ATMOSPHERIC PROPAGATION

-205-

For a plane wave input r'(p', 5 ,0) = JU1 2 the transmitter plane

integrals (A.18) can be performed and

rL ' L

r'(pc'PdL) = JR12 exp [- a(z)dz exp -[1-P(Pd 0 (z)dz]

If a Gaussian form for the phase function is assumed

exp(- Is 12 /2 2)

F

(A.20)

(A.21)

then P(pd) is

P(d)= exp(-k 2e .Id 2 /2)

(A.22)

1 1 - k2 e2 d 2 / 2

where a two-term Taylor series expansion of P has been indicated. If

we use this two-term Taylor series and further assume a uniform

scattering profile along the propagation path a(z) = S(z) = S'

then (A.19) becomes

<hL (1,p ) h*(p 2 ,p )> = ho(phpi) h (p2' 2

-e-aL expf Ed 2 + d d+ Id 2

3p(A.23)

Page 206: ATMOSPHERIC PROPAGATION

-206-

where

p2 2 (A.24)0 asL k2 e

which for plane wave input implies

- L -1 12 p

r'(PC'Pd,L) = JU 2 e e (A.25)

Equations (A.20) and (A.25) have been plotted in Figure A.3. It is

clear that if sL >> 1, then (A.25) is a good approximation to (A.20).

By extension we say that under the same condition (A.23) is a good

approximation to (A.19). In order to see what is neglected by using

(A.23) in place of (A.19) consider again Figure A.3. The correlation

that remains between the field sampled at two highly separated points

in the upper curve (for large IPd1, ' =u2 exp(-(a + S)L)) can

only be due to unscattered light as scattered light should, by

physical reasoning, become uncorrelated at large 1-d'. Therefore we

conclude that use of (A.25) in place of (A.20) and, by extension (A.23)

in place of (A.19), amounts to treating only the scattered light and

disregards the unscattered beam. Hence, the unscattered power needs to

be considered separately from the scattered power.

To conclude our development of the MFS propagation model

consider the sketch of a real phase function in Figure A.4. Besides

being highly peaked in the forward direction there is significant

Page 207: ATMOSPHERIC PROPAGATION

1'

Jul 2 e-pa L

1 2 - L

2 e~

-e3s

Ra L 2-P4 2

0

Figure A.3: MCF's for plane wave input.

2Iy I

-I

e-(lea + G) L

d

L (I- P(p )

Page 208: ATMOSPHERIC PROPAGATION

-208-

wide angle and back scatter. Our aim here is to apply the previous

theoretical development to such a real world situation. The procedure

we follow has previously been used by Ross et al. [18] and Mooradian

et al. [36]. Reasonably good, but inconclusive experimental

verification of this prcoedure has been reported by both groups of

researchers. First, we truncate the real phase function at eE, as

shown in Figure A.4, chosen to contain the forward scatter peak and

lump all scattering at angles wider than 8E into a modified absorption

coefficient. Hence we would use

; = Na + s(1 @ (A.26)

S = s (A.27)

In place of a, s in (A.23), (A.24) where

a+ s a + s t (A.28)

and

-e=2Tr p(8) sin e de (A.29)

-10

Page 209: ATMOSPHERIC PROPAGATION

-209-

4

90

8 (deg)

Real phase function.

p (e)

1808 GE

Figure A.4:

Page 210: ATMOSPHERIC PROPAGATION

-210-

is the forward scattering efficiency. Second, the forward scatter

peak is approximated by the Gaussian phase function (A.21) where 6F3

the effective rms forward scatter angle, is given by [17]

6 2-rr 0 (A.30)F 27rp(p)J

Clearly the forward scatter efficiency must satisfy 0 < 0 < 1. To

simplify our calculations in Chapter VII, we choose D = 0.57 as a

reasonable value [17] in place of (A.29). Equation (2.C.2) is therefore

(A.23) with parameters chosen according to the above procedure.

Page 211: ATMOSPHERIC PROPAGATION

-211-

REFERENCES

1. J.H. Shapiro, "Imaging and Optical Communication throughAtmospheric Turbulence," Topics in Applied Physics, Vol. 25:Laser Beam Prupdqatio. in the Atmosphere, Ed. J.W. Strohbehn.

2. J.H. Shapiro, "Extended Huygens-Fresnel Principle," InternalMemo, Optical Propagation and Communication Group, MIT,Cambridge, MA, March 1979.

3. J.H. Shapiro, "Imaging and Target Detection with a Heterodyne-Reception Optical Radar," Project Report TST-24, LincolnLaboratory, MIT, October 1978.

4. B.A. Capron, R.C. Harney, and J.H. Shapiro, "Turbulence Effectson the Receiver Operating Characteristics of a Heterodyne-Reception Optical Radar," Project Report TST-33, LincolnLaboratory, MIT, July 1979.

5. J.H. Shapiro, B.A. Capron, and R.C. Harney, "Imaging and TargetDetection with a Heterodyne-Reception Optical Radar,"Appi. Opt., Vol. 20, No. 19, Oct. 1, 1981, pp. 3292-3313.

6. R.C. Harney and R.J. Hull, "Compact Infrared Radar Technology,"Proc. SPIE, Vol. 227, pp. 162-70, 1980.

7. R.J. Becherer, "Pulsed Laser Ranging Techniques at 1.06 and10.6 -Wm," Project Report TT-8, Lincoln Laboratory, MIT,March 1976.

8. R.J. Becherer, "System Design Study for Infrared Airborne Radar(IRAR)," Technical Note 1977-29, Lincoln Laboratory, MIT,October 1977.

9. J.W. Goodman, Introduction to Fourier Optics, McGraw Hill, NewYork, 1968, pp. 57-61.

10. J.H. Shapiro, "Reciprocity of the Turbulent Atmosphere," J. Opt.Soc. Am. 61, 492-495 (1971).

11. P. Bogler, "Reciprocity for Propagation in a Turbid Media,"Private Communication.

12. A.E. Siegman, "The Antenna Properties of Optical HeterodyneReceivers," Proc. IEEE 54, pp. 1350-56 (1966).

Page 212: ATMOSPHERIC PROPAGATION

-212-

13. R.M. Gagliardi and S. Karp, Optical Communications, John Wileyand Sons, New York, 1976, pp. 173-81.

14. W.B. Davenport, Probability and Random Processes, McGraw-Hill,New York, 1970, pp. 499-500.

15. W.B. Davenport and W.L. Root, An Introduction to the Theory ofRandom Signals and Noise, McGraw-Hill, New York, 1958,pp. 81-84.

16. D.L. Fried, "Aperture Averaging of Scintillation," J. Opt. Soc.Am. 57, 169-175, 1967.

17. J. Nakai, "A Spatio-Temporal Channel Model of Low-VisibilityOptical Wave Propagation," Masters Thesis, Dept. of EE&CS,MIT, January 1980.

18. W.S. Ross, W.P. Jaeger, J. Nakai, T.T. Nguyen and J.H. Shapiro,"Atmospheric Optical Propagation: An Integrated Approach,"to be published in Optical Engineering.

19. J.W. Goodman, "Some Effects of Target-Induced Scintillation onOptical Radar Performance," Proc. IEEE 53, 1688-1700, 1965.

20. J.C. Dainty, ed.; Laser Speckle and Related Phenomena (Springer-Verlag, Berlin, 1975).

21. D.L. Fried, "Statistics of the Laser Radar Cross Section of aRandomly Rough Target," J. Opt. Soc. Am. 66, 1150-1160,1976.

22. Target Signature Analysis Center: Data Compilation EleventhSupplement, Vol. 1 - Bidirectional Reflectance: Definition,Discussion and Utilization and Vol. 2 - BidirectionalReflectance: Graphic Data, AFAL-TR-72-226, (1972).

23. H.L. Van Trees, Detection Estimation and Modulation Theory,Vol. 1, J. Wiley and Sons, New York, 1968.

24. J.H. Shapiro and C. Warde, "Optical Communication through LowVisibility Weather," Optical Engineering, January, February1981, Vol. 20, No. 1, pp. 76-83.

25. A. Ishimaru, Wave Propagation and Scattering in Random Media,Vol. 2, Academic, New York, pp. 296-301, 1978.

26. D.M. Papurt, J.H. Shapiro, and R.C. Harney, "AtmosphericPropagation Effects on Coherent Laser Radars," to appearin Proceedings of the SPIE, Vol. 300.

Page 213: ATMOSPHERIC PROPAGATION

-213-

27. J.H. Shapiro and S.T. Lau, "Turbulence Effects on CoherentLaser Radar Target Statistics," to appear in AppliedOptics.

28. A. Ishimaru, Wave Propagation and Scattering in Random Media,Vol. 1, Academic, New York, 1978.

29. C. de Boor, "CADRE: An Algorithm for Numerical Quadrature,"Mathematical Software, (John R. Rice, ed.) AcademicPress, New York, 1971, Chapter 7.

30. I.S. Gradshteyn and I.M. Ryzhik, Table of Integrals, Series andProducts, Academic Press, New York, 1980.

31. J.H. Shapiro, "MIT Target Detection Analysis," Internal Memo,Lincoln Laboratories, Lexington, MA, December 29, 1978.

32. H.L. Van Trees, Detection, Estimation and Modulation Theory,Vol. 1, J. Wiley and Sons, New York, 1968, Sect. 2.7.

33. S.J. Halme, B.K. Levitt and R.S. Orr, "Bounds and Approximationsfor Some Integral Expressions Involving LognormalStatistics," Res. Lab. Electron. Quart. Prog. Rept. 93,163-175, MIT (1969).

34. E.V. Hoversten, R.O. Harger and S.J. Halme, "CommunicationTheory for the Turbulent Atmosphere," Proc. IEEE 58,1626-1650 (1970).

35. S.T. Lau, "Pulse-Integration Detection Performance," InternalMemo, Lincoln Laboratories, Lexington, MA, July 23, 1981.

36. G.C. Mooradian, M. Geller, L.B. Stotts, D.H. Stephens, andR.A. Krautwald, "Blue-Green Pulsed Propagation ThroughFog," Applied Optics, Vol. 18, No. 4, pp. 429-441, February1979.

37. D. Deirmendjian, Electromagnetic Scattering on SphericalPolydispersions, American Elsevier, 1969.

38. R.L. Anderson and T.A. Bancroft, Statistical Theory in Research,McGraw-Hill, 1952.

39. C. Chatfield, Statistics for Technology, Halsted Press, 1970.

40. T.S. Chu and D.C. Hogg, "Effects of Preciptation on Propagationat .63, 3.5 and 10.6 Microns," Bell System TechnicalJournal, May-June 1968, pp. 723-759.

Page 214: ATMOSPHERIC PROPAGATION

-214-

41. R.S. Lawrence and J.W. Strohbehn, "A Survey of Clear-AirPropagation Effects Relevent to Optical Communications,"Proc. IEEE 58, 1523-1545, 1970.

Page 215: ATMOSPHERIC PROPAGATION

-215-

BIOGRAPHICAL NOTE

David Papurt was born in Toledo, Ohio on July 18, 1954.

He graduated from Thomas A. DeVilbiss High School in June 1972. Dr.

Papurt then entered the University of Toledo where he received the

B.S. degree in Electrical Engineering in June 1977 and the B.A.

degree in Music in August 1977.

In September 1977, Dr. Papurt became a graduate student in

the Department of Electrical Engineering and Computer Science at

the Massachusetts Institute of Technology where he specialized in

communication theory. He received the S.M. degree in Electrical

Engineering and Computer Science and the E.E. degree from MIT in

September 1979 and June 1980, respectively. He is a member of the

Society of Photo-Optical Instrumentation Engineers.

Dr. Papurt is currently Assistant Professor of Electrical

Engineering at Northeastern University, Boston, Massachusetts. He

has co-authored a paper on "Atmospheric Propagation Effects on

Coherent Laser Radars," that will appear in the Proceedings of the

Society of Photo-Optical Instrumentation Engineers.