Remote sensing of sediment characteristics by optimized echo ...

17
Remote sensing of sediment characteristics by optimized echo-envelope matching a) Daniel D. Sternlicht b) and Christian P. de Moustier c) Marine Physical Laboratory, Scripps Institution of Oceanography, University of California at San Diego, La Jolla, California 92093-0205 ~Received 1 November 1999; revised 13 June 2003; accepted 26 June 2003! A sediment geoacoustic parameter estimation technique is described which compares bottom returns, measured by a calibrated monostatic sonar oriented within 15° of vertical and having a 10°–21° beamwidth, with an echo envelope model based on high-frequency ~10–100 kHz! incoherent backscatter theory and sediment properties such as: mean grain size, strength, and exponent of the power law characterizing the interface roughness energy density spectrum, and volume scattering coefficient. An average echo envelope matching procedure iterates on the reflection coefficient to match the peak echo amplitude and separate coarse from fine-grain sediments, followed by a global optimization using a combination of simulated annealing and downhill simplex searches over mean grain size, interface roughness spectral strength, and sediment volume scattering coefficient. Error analyses using Monte Carlo simulations validate this optimization procedure. Moderate frequencies ~33 kHz! and orientations normal with the interface are best suited for this application. Distinction between sands and fine-grain sediments is demonstrated based on acoustic estimation of mean grain size alone. The creation of feature vectors from estimates of mean grain size and interface roughness spectral strength shows promise for intraclass separation of silt and clay. The correlation between estimated parameters is consistent with what is observed in situ. © 2003 Acoustical Society of America. @DOI: 10.1121/1.1608019# PACS numbers: 43.30.Gv, 43.30.Hw, 43.30.Ft, 43.30.Pc @DLB# Pages: 2727–2743 I. INTRODUCTION Remote classification of ocean sediments is motivated by mineral resources assessment, cable and pipeline route planning, and mine warfare. In recent years a number of high-frequency ~.10 kHz! echo analysis techniques have been developed for characterizing the upper layer of seafloor sediments. Sediment classification techniques using single-beam so- nars are either phenomenological or physical. Phenomeno- logical approaches identify nonparametric measured echo characteristics with core samples or bottom photographs. Such systems typically require calibration of signal charac- teristics with ground truth at the beginning of each survey, and operation must proceed at a fixed sensor altitude. Pace and Ceen investigated sediment characterization using single-beam echoes, 1 where comparison of the expanded echo ~due to temporal spreading! with the transmit pulse was used to infer bottom roughness. Echo durations commensu- rate with the duration of the transmit pulse were thought to originate from smooth substrates, whereas longer, variably shaped echoes were attributed to coarse materials. Sediment classification techniques that empirically match echo charac- teristics to ground truth have since been developed. One such system 2 exploits the bottom echo and the first surface mul- tiple ~bottom–surface–bottom! by integrating the energy over the tail section of the first return, and integrating over the entire length of the multiple. Representation of these two measures as feature vectors allows segregation of a variety of bottom types. Building on this paradigm, multifeature clas- sification techniques based on higher moment statistics of the recorded waveform are being investigated. 3,4 Results from Ref. 1 inspired interpretation of the bottom echo’s tail as an indicator of bottom roughness, while the energy content of the multiple is considered an indicator of the reflection coefficient, or hardness of the substrate. Theo- retical explanations for the success of these systems and modeling of the bistatic geometry are being investigated. 5,6 In physics-based approaches, sediment characteristics are estimated by comparing measurements to predictions made with physical models—thus minimizing presurvey training requirements and removing limitations on sensor altitude that, typically, are found in phenomenological ap- proaches. One example of physics-based acoustic sediment characterization is described in the works of Schock, LeBlanc, and Mayer, 7,8 wherein broadband ~2–10 kHz! echo amplitudes are used to estimate coherent reflection coeffi- cients of sediment layers, and measured distortions of echo spectra yield information on sediment attenuation properties. The inspiration for our work comes from physics-based echo envelope inversion techniques described by Berry, 9 Nesbitt, 10 Jackson and Nesbitt, 11 and Lurton and Pouliquen. 12 Berry’s estimation of irradiated surface charac- teristics employs half-power lengths of measured and mod- eled average radar backscatter envelopes. Nesbitt used a least-squares search for matching acoustic backscatter enve- lopes with models based on reflection loss, sediment absorp- tion coefficient, rms bottom slope, and a sediment volume scattering parameter. His work incorporated up to two sedi- a! Parts of this manuscript were presented in the talks: Sternlicht and de Moustier @J. Acoust. Soc. Am. 105, 1206 ~1999!# and Sternlicht and de Moustier @J. Acoust. Soc. Am. 108, 2536 ~2000!#. b! Current address: Dynamics Technology Inc., 21311 Hawthorne Blvd., Suite 300, Torrance, CA 90503. Electronic mail: [email protected] c! Current address: Center for Coastal and Ocean Mapping, University of New Hampshire, 24 Colovos Road, Durham, NH 03824. Electronic mail: [email protected] 2727 J. Acoust. Soc. Am. 114 (5), November 2003 0001-4966/2003/114(5)/2727/17/$19.00 © 2003 Acoustical Society of America

Transcript of Remote sensing of sediment characteristics by optimized echo ...

Page 1: Remote sensing of sediment characteristics by optimized echo ...

Remote sensing of sediment characteristics by optimizedecho-envelope matchinga)

Daniel D. Sternlichtb) and Christian P. de Moustierc)

Marine Physical Laboratory, Scripps Institution of Oceanography, University of California at San Diego,La Jolla, California 92093-0205

~Received 1 November 1999; revised 13 June 2003; accepted 26 June 2003!

A sediment geoacoustic parameter estimation technique is described which compares bottomreturns, measured by a calibrated monostatic sonar oriented within 15° of vertical and having a10°–21° beamwidth, with an echo envelope model based on high-frequency~10–100 kHz!incoherent backscatter theory and sediment properties such as: mean grain size, strength, andexponent of the power law characterizing the interface roughness energy density spectrum, andvolume scattering coefficient. An average echo envelope matching procedure iterates on thereflection coefficient to match the peak echo amplitude and separate coarse from fine-grainsediments, followed by a global optimization using a combination of simulated annealing anddownhill simplex searches over mean grain size, interface roughness spectral strength, and sedimentvolume scattering coefficient. Error analyses using Monte Carlo simulations validate thisoptimization procedure. Moderate frequencies~33 kHz! and orientations normal with the interfaceare best suited for this application. Distinction between sands and fine-grain sediments isdemonstrated based on acoustic estimation of mean grain size alone. The creation of feature vectorsfrom estimates of mean grain size and interface roughness spectral strength shows promise forintraclass separation of silt and clay. The correlation between estimated parameters is consistentwith what is observedin situ. © 2003 Acoustical Society of America.@DOI: 10.1121/1.1608019#

PACS numbers: 43.30.Gv, 43.30.Hw, 43.30.Ft, 43.30.Pc@DLB# Pages: 2727–2743

teoo

eo

snchh

aceya

sind

nstb

mrasuul

ertwoty ofs-the

mtheofeo-and

sticsionseysorp-

mentck,

effi-choies.edry,

c-od-ed anve-orp-medi-

vdo

yai

I. INTRODUCTION

Remote classification of ocean sediments is motivaby mineral resources assessment, cable and pipeline rplanning, and mine warfare. In recent years a numberhigh-frequency~.10 kHz! echo analysis techniques havbeen developed for characterizing the upper layer of seaflsediments.

Sediment classification techniques using single-beamnars are either phenomenological or physical. Phenomelogical approaches identify nonparametric measured echaracteristics with core samples or bottom photograpSuch systems typically require calibration of signal charteristics with ground truth at the beginning of each survand operation must proceed at a fixed sensor altitude. Pand Ceen investigated sediment characterization usingle-beam echoes,1 where comparison of the expandeecho~due to temporal spreading! with the transmit pulse wasused to infer bottom roughness. Echo durations commerate with the duration of the transmit pulse were thoughtoriginate from smooth substrates, whereas longer, variashaped echoes were attributed to coarse materials. Sediclassification techniques that empirically match echo chateristics to ground truth have since been developed. Onesystem2 exploits the bottom echo and the first surface mtiple ~bottom–surface–bottom! by integrating the energy

a!Parts of this manuscript were presented in the talks: Sternlicht andMoustier @J. Acoust. Soc. Am.105, 1206 ~1999!# and Sternlicht and deMoustier @J. Acoust. Soc. Am.108, 2536~2000!#.

b!Current address: Dynamics Technology Inc., 21311 Hawthorne BlSuite 300, Torrance, CA 90503. Electronic mail: [email protected]

c!Current address: Center for Coastal and Ocean Mapping, UniversitNew Hampshire, 24 Colovos Road, Durham, NH 03824. Electronic [email protected]

J. Acoust. Soc. Am. 114 (5), November 2003 0001-4966/2003/114(5)/2

dutef

or

o-o-o

s.-,ceg

u-olyentc-ch-

over the tail section of the first return, and integrating ovthe entire length of the multiple. Representation of thesemeasures as feature vectors allows segregation of a variebottom types. Building on this paradigm, multifeature clasification techniques based on higher moment statistics ofrecorded waveform are being investigated.3,4

Results from Ref. 1 inspired interpretation of the bottoecho’s tail as an indicator of bottom roughness, whileenergy content of the multiple is considered an indicatorthe reflection coefficient, or hardness of the substrate. Thretical explanations for the success of these systemsmodeling of the bistatic geometry are being investigated.5,6

In physics-based approaches, sediment characteriare estimated by comparing measurements to predictmade with physical models—thus minimizing presurvtraining requirements and removing limitations on senaltitude that, typically, are found in phenomenological aproaches. One example of physics-based acoustic sedicharacterization is described in the works of SchoLeBlanc, and Mayer,7,8 wherein broadband~2–10 kHz! echoamplitudes are used to estimate coherent reflection cocients of sediment layers, and measured distortions of espectra yield information on sediment attenuation propert

The inspiration for our work comes from physics-basecho envelope inversion techniques described by Ber9

Nesbitt,10 Jackson and Nesbitt,11 and Lurton andPouliquen.12 Berry’s estimation of irradiated surface charateristics employs half-power lengths of measured and meled average radar backscatter envelopes. Nesbitt usleast-squares search for matching acoustic backscatter elopes with models based on reflection loss, sediment abstion coefficient, rms bottom slope, and a sediment voluscattering parameter. His work incorporated up to two se

de

.,mofl:

2727727/17/$19.00 © 2003 Acoustical Society of America

Page 2: Remote sensing of sediment characteristics by optimized echo ...

ctoump

deeden

grrmrin

itrecos

epdiatht

t

tio-d

e

o

nlie

th

rua

um

ts

t

encotoayop

ooreamby

-ten-tterwasli-

n tode-

de-

ateam

gleicalres-lopesage,s on-an-g.mitfil-

rednt at’singam

ics-ents,ionhond

urenti-

therac-r-tedvednceionl

ta,e-ceroure

coustic backscatter models for extracting bottom charaistics from single-beam echo-sounder data. Lurton and Piquen described a method for sea-bottom identification coparing normalized cumulative functions of the echo enveloderived from measurements and physical backscatter moWaveform normalization allows for the use of uncalibratecho-sounders; however, ignoring echo strength limitsploitation of important information such as impedance cotrast at the water–sediment interface. Furthermore, intetion of echo envelope time series into cumulative fodisproportionately represents signal components occurearlier in time~closer to normal incidence!.

In this work, we match a physics-based echo intensenvelope model13 to seafloor acoustic backscatter measuments, made over substrates ranging from clay to sand,lected with calibrated 33- and 93-kHz echo-sounders wh3-dB beamwidths~10°–21°! and elevation angles~maximumresponse axis at 0°–15° incidence! are consistent with themodel’s underlying Kirchhoff scattering theory. This modincorporates the system’s deployment geometry, beamtern, and signal characteristics, the ocean volume spreaand absorption losses, and solutions of the monochromwave equation using boundary conditions described bysediment geoacoustic characteristics. The time-dependentensity measured at the transducer faceI (t) is modeled as thesum of a sediment interface componentI i(t) and a sedimenvolume componentI v(t)

I ~ t !5I i~ t !1I v~ t !, ~1!

where, following the theoretical work of Jacksonet al.,14 theinterface backscatter component is obtained from a soluof the Helmholtz diffraction integral using the Kirchhoff approximation, and a composite roughness approach is usepredict scattering from the sediment volume.

Model parameters include the mean grain size (Mf),defined asMf52 log2 Dg , whereDg is the sediment’s meangrain diameter,15,16 and its correlates, the sediment:watdensity and sound-speed ratios~r,n!, and the sediment’scompressional wave attenuation constant (kp indB/m/kHz!.17 Fluctuations of these properties are incorprated into a sediment scattering coefficient,sv (m21), signi-fying the scattering cross section per unit volume, per usolid angle. The interface is modeled by a power-law reenergy density spectrumW(k)5w2k2g, wherek is the bot-tom relief’s two-dimensional wave number vector wimagnitudek, w2 is the spectral strength~expressed in unitscm4!, andg is the spectral exponent. The roughness spectis bandlimited to wave numbers spanning approximatelyorder of magnitude above and below the acoustic wave nber.

The expectedin situ ranges of the model componenare: 21<Mf<9, 2.4<g<3.9, 0.0<w2<1.0, 0.8<n<3.0, 1.0<r<3.0, 0.01<kp<1, 0.0<sv<1.0ab. ab

is the sediment compressional wave attenuation coefficiendB/m, calculated asab5kp3 f a ,18 and f a is the acousticfrequency in kHz. If the statistics describing the sedimcharacteristics are consistent over measurement scalesmensurate with the geographic range of collected botechoes, the geoacoustic parameters described above mestimated from optimized comparisons of the echo envel

2728 J. Acoust. Soc. Am., Vol. 114, No. 5, November 2003

erl--els.

x--a-

g

y-

ol-e

lat-ngticein-

n

to

r

-

itf

mn-

in

tm-

mbee

model with aligned and averaged data.Normalized angular dependence curves of seafl

acoustic backscatter, measured with the 16-beam SeaBecho-sounder, fitted with computed curves parametrizedthe relief spectrum components (g,w2) was presented byMichalopoulouet al.19 Using a least-squares maximum likelihood estimator and chi-square acoustic backscatter insity statistics, the potential of matching acoustic backscamodels with statistically independent measurementsdemonstrated. A drawback of this implementation is its reance on exhaustive search procedures and its limitatiohigh-impedance contrast, impenetrable substrates with nomonstrable volume component. Another approach isscribed in Matsumotoet al.,20 where global optimization bysimulated annealing and downhill simplex is used to estimrelief spectrum parameters from the same kind of SeaBeacoustic data.

The samples of the time series measured with a sinecho-sounder are partially correlated, making the statistapproach of Ref. 19 inappropriate. Instead, the model’s psure time series are matched to measured echo envecalculated from stacked and averaged data with a two-staverage echo envelope matching procedure, which buildthe work of Matsumotoet al.20 by expanding the optimization to include relief spectrum parameters and physical qutities related to grain size and sediment volume scatterin

By incorporation of the measurement system’s transand receive sensitivities, directional characteristics, and atering operation for converting voltage waveforms measuat the transducer terminals to pressure waveforms incidethe transducer,21 the shape and amplitude of the bottomangular response is exploited in a model–data matchscheme appropriate for simple, inexpensive, single-beecho sounders. This is distinguished from other physbased approaches which compare normalized measuremof uncalibrated returns to normalized model realizations12

and from phenomenological seafloor characterizattechniques2,4 using correlation analysis of measured ecfeatures~e.g., amplitude and energy in bottom echoes arespective surface multiples! with known ground truth.

Our physics-based model–data optimization procedgenerates feature vectors with elements consisting of quafiable geoacoustic parameters (Mf ,w2 ,sv). This informa-tion can be directly associated with bottom type; thus,procedure is, in theory, independent of specific site chateristics or insonification geometry~such as water depth otransducer orientation!. In addition, the sensitivity of this optimization procedure to echo variability can be estimafrom the covariance matrix of geoacoustic features, derifrom synthetic data sets generated with the data covariamatrix for an ensemble of returns. Furthermore, correlatbetween the geoacoustic parameters~whether due to naturaphenomena or artifacts of the optimization procedure! can becharacterized.

Computation of the average echo envelope from daand of a signal to error ratio in the model–data fit are dscribed in Sec. II, with an example of the data covarianmatrix and its implications to the model–data matching pcedure. The two-stage model–data optimization proced

D. D. Sternlicht and C. P. de Moustier: Remote sensing of sediment

Page 3: Remote sensing of sediment characteristics by optimized echo ...

IIt trictepi

secn

o

sio

aeninch

bc

ragplTdr

forwayiortivecho

ingdataal

orali-

for estimating bottom characteristics is presented in Sec.with a method of evaluating the error propagation inherenthe matching procedure, using parameter covariance matproduced from Monte Carlo simulated data sets. The sysand data used to validate the echo envelope model andrameter estimation technique are described in Sec. IV, wresults presented in Sec. V, and analyses of the effectecho variability on the optimization procedure given in SVI. Section VII draws conclusions about the usefulness ashortcomings of the approach, and its potential for seaflclassification.

II. AVERAGE ECHO ENVELOPE

The measured bottom echo consists of a pulsed CWnal, modulated by the bottom backscattering process, whenvelope detection and sampling at periodte yield an rmspressure sequence,p@n#, expressed in units of pascals~Pa!.Acoustic wavelengths at frequencies greater than 10 kHzgenerally small compared to the relief of the water–sediminterface, and bottom echoes are incoherent, varying sigcantly in amplitude and shape as the sonar translates lotudinally above the interface. Because of this variability, eoes must be treated stochastically.

For comparison with the temporal model, an ensemof M contiguous returns is characterized by the average esequence (pa@n#, n50,1,...,N21). To this end, a two-dimensional amplitude arrayp@m,n# is defined for (0<n<N21) samples per ping and (0<m<M21) pings, incor-porating segments of the data presented in Sec. IV

pa@n#51

M (m50

M21

p@m,n#, n50,1,...,N21. ~2!

Samples in the echo envelopes from the incoherentturns are Rayleigh distributed, but their ensemble averover many pings is approximately Gaussian. Hence, samof the average echo envelope are Gaussian distributed.N3N covariance matrixC of the average echo is estimateby normalizing the data sample covariance by the numbereturns~M!. Elements ofC are thus

FIG. 1. Average echo envelope for silt substrate:f a533 kHz, maximumresponse axis at 8° incidence. Solid line ispa@n# @Eq. ~2!#, dashed lines arepa@n#6sa@n#.

J. Acoust. Soc. Am., Vol. 114, No. 5, November 2003 D. D.

I,oesma-

thof.d

or

g-se

rentfi-gi--

leho

e-e

eshe

of

Ci j 51

M H 1

M21 (m50

M21

~p@m,i #2pa@ i # !~p@m, j #2pa@ j # !J ,

~3!

where (0< i , j <N21). Henceforth,C is referred to as thedata covariance matrix.

To focus this description, we use the average echo100 consecutive returns measured from a vessel underover a silt substrate in San Diego Bay, plotted in Fig. 1. Prto averaging, the echoes were aligned along their respecthreshold indices as described in Ref. 13. The average eenvelope is bracketed bypa@n#6sa@n#, where variancessa

2@n# correspond to the diagonal elements ofC. Plots ofCand its corresponding correlation coefficient matrixY, withelements:Y i j 5Ci j /sa@ i #sa@ j # ~Fig. 2! show that the vari-ance is proportional to signal strength and that neighborsamples are highly correlated. In later sections, syntheticsets generated withC will help assess the effects of signvariability on the model–data matching procedure.

The average echo is summarily matched by a tempmodel estimate (pa@n#) generated with specified mean alttude and sediment geoacoustic parameters

FIG. 2. ~a! Data covariance matrix@Eq. ~3!#, and~b! Correlation coefficientmatrix, for average echo envelope of Fig. 1.

2729Sternlicht and C. P. de Moustier: Remote sensing of sediment

Page 4: Remote sensing of sediment characteristics by optimized echo ...

thr

ehn

hssanbe

migomt

raa

umchfi

inc

eri-

thec

andtedel–. Iteserch-ncethef arch

al-

g

ce,er-tad by

r-

aetera-

chpro-oa-s ofrchobal

rao

etetehi er

pa@n#5ArwnwI a@n#, ~4!

where I a@n# represents the discrete implementation ofecho envelope model, andrw andnw correspond to seawatedensity and sound speed, respectively.

To measure the fit between the model and data, a mfunction compares the total energy in the average ecpa@n#, to a measure of energy representing the discrepabetween model and data. This signal to error ratio~S/E! isexpressed as

S/E5(n5n1

n2 pa2@n#

(n5n1

n2 ~pa@n#2 pa@n# !2, ~5!

where n1 and n2 are the initial and final indices for botwaveforms. In this scheme a high value of S/E signifie‘‘good’’ match of model with data. This method providenumerical evaluations which are independent of scalesignal length, and is convenient for comparing resultstween data sets.

III. GEOACOUSTIC PARAMETER ESTIMATION

The estimation of bottom characteristics from the teporal model depends on a model–data matching parad~Fig. 3! that converges to a unique and correct set of bottparameters. The bottom characteristics which describedata are determined by comparing the model to the avebottom echo, with the goal of minimizing the error to signratio ~E/S!, i.e., the inverse of Eq.~5!. However, estimationof geoacoustic parameters is complicated by the large nber of good fits existing in the multidimensional searspace, where it is possible to find convincing model–datawhich do not necessarily represent correct solutions.13 Arriv-ing at sensible solutions requires parsing the problemmanageable parts, establishing the degree of parameterrelations, and constraining the search space.

A. Two-stage parametric optimization

With the goal of deriving unambiguous matches btween the temporal model and data, we initially expemented with a one-dimensional~1D! search technique, ex

FIG. 3. Geoacoustic parameter optimization procedure: The compafeeds back~E/S! to the parameter selection module to guide the selectionmore promising parameter settings. The system outputs model paramcorresponding to the optimal fit. Careful implementation of the paramselection module determines the success and tractability of this matcprocedure.

2730 J. Acoust. Soc. Am., Vol. 114, No. 5, November 2003

e

rito,cy

a

d-

-m

hegel

-

ts

toor-

--

tracting the best-fit generic characteristics by iterating onmean grain size parameter (Mf). Here, the six geoacoustiparameters (w2 ,g,sv ,r,n,kp) are related toMf throughlinear regression formulas adapted from Refs. 17, 18summarized in Appendix A of Ref. 13. As was demonstrain Ref. 21, the generic parameters produce rough moddata fits for the San Diego Bay substrates investigatedfollows that the solution produced with the 1D search defina seed vector (Mf ,w2 ,sv) appropriate for a second-stagmultiparameter search in whichg is held to a constant. Fothe second stage, multiparameter local optimization teniques yielded disappointing results marked by convergeto solutions which were unstable and overly sensitive tochoice of seed vector. This led to the development omodel–data matching procedure incorporating the 1D seato establish the general sediment type~sand or fines! and thespectral exponent~g!, followed by a three-dimensional~3D!global optimization using a combination of simulated anneing and downhill simplex searches~SA/DS! over the rough-ness spectral strength (w2), the sediment volume scatterincoefficient (sv), and the mean grain size (Mf) associatedwith the correlated parameters:r, n, kp .

1. Stage 1: 1D golden section search and parabolicinterpolation

For transducer orientations close to normal incidenthe bottom reflection coefficient is the dominant factor detmining the signal amplitude. It follows that the model vs dasearch space generally has one extremum when describethe single parameterMf . This situation is illustrated by theE/S vs Mf plot of Fig. 4~a!, where the ‘‘best’’ solution isfound by iteratively bracketing the minimum. For this pupose, we employ a combination of thegolden sectionsearchalgorithm coupled with inverse parabolic interpolation,procedure formulated in Ref. 22. The geoacoustic paramoutputs of stage 1 provide a starting point for the multiprameter global search technique of stage 2.

2. Stage 2: Global simulated annealing—downhillsimplex optimization (SA ÕDS)

After testing a number of local multiparameter seartechniques, we found nongreedy, nonexhaustive searchcedures to be most appropriate for finding the best-fit gecoustic parameters. These techniques investigate regionthe parameter space not typically visited by local seatechniques, thus increasing the prospects that a true gl

torfersr

ngFIG. 4. Parameter space representations for optimization algorithms:~a! 1Dsearch space, E/S vsMf ; ~b! Reflectionacross the face of a three parametsimplex.

D. D. Sternlicht and C. P. de Moustier: Remote sensing of sediment

Page 5: Remote sensing of sediment characteristics by optimized echo ...

lype

onneee-aosis

rathc

seo

lor–mib

am-

theonse of

i-gi-

ionout

orallo-

acecsueam-am-ig.theere

-

al

ec-Therate

rre-

ondan

f they ism-

rical

eytede

o-his

ceex-subto

minimum will be found. Insimulated annealingthe systemis initialized to some high-energy state and then slowbrought to the zero state, where a final local search isformed.

A variety of annealing techniques exists, with commreliance on randomly generated numbers for selection ofparameter vectors. We initially tested the best-known vsion, described in Ref. 23. This method employs the Mtropolis algorithm,24 for which randomly generated parameter vectors, yielding a lower cost than the current vector,automatically accepted, while those yielding a higher care accepted by condition of the Boltzmann probability dtribution

P~DE!5exp~2DE/T!, ~6!

whereDE signifies a positive increase in energy at tempeture T. If the search space is vast and/or if calculation ofobjective function is computationally intensive, convergenfor this method may be unacceptably slow.

Although the temporal model lacks analytic derivativeit is continuous in the sense that a small change in paramvalue is accompanied by a proportional change in the cfunction. With this information, faster convergence to a gbal minimum may be achieved by employing the NeldeMead downhill simplex search, modified by randotemperature-dependent uphill energy transitions as descrin Ref. 25. For a solution space comprised of three pareters (Mf ,w2 ,sv), a simplex of four solution vectors is cre

FIG. 5. Flow chart for parameter estimation.

J. Acoust. Soc. Am., Vol. 114, No. 5, November 2003 D. D.

-r-

wr--

ret-

-ee

,terst-

ed-

ated as illustrated in Fig. 4~b!. The cost function E/S is mini-mized by reflections, contractions, and expansions ofsimplex where, at high temperatures, nonoptimal solutiare occasionally accepted into the simplex at the expensbetter solutions. At the final temperature stage (T50) thesimplex is assumed to be in the vicinity of the global minmum, and the Nelder–Mead algorithm is applied in its orinal form, only accepting better solutions~local search!. Tomaximize the algorithm’s effectiveness, the best solutfound since initiation of the search is preserved throughthe annealing process.

B. Parameter estimation paradigm

The geoacoustic parameters contained in the tempmodel define a complicated search space with numerouscal minima. It is thus essential to constrain the solution spusinga priori knowledge, and to employ practical heuristiin order to reject implausible solutions. For extracting uniqand meaningful sediment parameters from the shape andplitude of measured bottom echoes, we propose the pareter estimation paradigm illustrated by the flow chart of F5. This technique represents an automated version ofmodel–data matching guidelines proposed in Ref. 13, whthe result of the initial 1D local search~top module of theflow chart! provides thea priori information needed to constrain the second stage. TheMf result is fed to a decisionjunction which determines the general bottom type~sands orfines! and sets the roughness spectral exponent~g! in prepa-ration for the multiparameter optimization. The 3D globSA/DS procedure iterates over a limited range ofMf , w2 ,andsv , fine-tuning the impedance contrast, roughness sptral strength, and volume estimates for the substrate.final result of this procedure provides the general substtype ~sand vs fines!, bottom characteristics (Mf ,g,w2 ,sv),and, indirectly, the sediment geoacoustic parameters colated to mean grain size.

It should be noted that the search space for the secstage optimization is constrained by restricting the megrain size to (Mf21)<Mf<(Mf11), whereMf repre-sents the seed value from stage one. When contortion oSA/DS simplex violates these bounds, a suitable penaltadded to the E/S cost function to reject out-of-bound paraeter vectors. Broad bounds are similarly applied to thew2

search space to avoid values unsuitable for the numeintegrations carried out by the temporal model algorithm.

The most important condition imposed on the volumscattering coefficient is (sv>0). However, unreasonabllarge volume components occasionally produce simulaechoes exhibiting low E/S scores. If the maximum volumcomponent is within 2 dB of the maximum interface compnent, an empirical penalty, proportional to the severity of tviolation, is added to the E/S cost function

IF I v /I i.0.63

THEN E/S5E/S* 4* H 115* S I v

I i20.63D J , ~7!

whereI v andI i represent the maximum volume and interfaintensities, respectively. This is a reasonable restrictioncept for oblique incidence measurements over fine-grainstrates, where it is possible for the volume component

2731Sternlicht and C. P. de Moustier: Remote sensing of sediment

Page 6: Remote sensing of sediment characteristics by optimized echo ...

onraith

hisrvthrabm

es

rb

rx

vahinatng

gar-

the

nd

-

-ratesingdis-

ofas

onsions

o-

ns

le,er-

1v

eral p

m-

dominate. In theory, the simulated annealing algorithm cverges asymptotically to an optimal solution if the tempeture is initially high and allowed to decrease inverse logarmically with the number of iterations.26 However, thecomputational requirements of the cost function in tmodel–data matching application require a more consetive number of model iterations. After experimenting withe annealing control parameters, adequate solution accuand convergence speed were achieved by employing a hylinear-exponential cooling schedule with nine discrete teperature levels. In this scheme, the initial temperatureT0 isset to the average E/S for the four initial simplex verticwhere one of these vectors~the seed! is derived from the firstlocal-search stage, and the other three are slightly pertureplicas. Ten model iterations are initially investigated atT0

and, for each temperature stage thereafter, the numbeiterations increases by 25%, resulting in a total of appromately 230 model iterations~e.g., Fig. 6!.

C. Evaluation of error propagation by Monte Carlosimulation

For a given bottom substrate, the average echo canfrom data ensemble to data ensemble. To characterizethis variation affects the results of the model–data matchprocedure,K synthetic average echo envelopes are generwith random combinations of signal and noise. Lacki

FIG. 6. Annealing process for the data shown in Fig. 1.w2 in cm4, sv inm21. The parameter values for the first iteration are the output of theoptimization; the final values are the annealing outputs. Note the largeriety of nonoptimal solutions investigated before low annealing temptures constrain the search space. In this particular example, the initiarameters are reasonably close to the final solution.

2732 J. Acoust. Soc. Am., Vol. 114, No. 5, November 2003

---

a-

cyrid-

,

ed

ofi-

ryowged

knowledge of the ‘‘true’’ signal, the model output resultinfrom the optimization procedure is distorted by noise chacterized in the data’s covariance matrix~C!. The optimiza-tion procedure is applied to each waveform and, usingresultingK solution vectors (Mf ,w2 ,sv), an approximationto the 333 parameter covariance matrix is computed aevaluated.

With bold lower case letters used to indicate 13N vec-tors, a simulated average echo envelope (ps) is calculatedfrom: the model outputpa, a vector of standard normal random deviates~x!, and the upper triangular matrix~A! fromCholesky factorization of the covariance matrix,C5ATA

ps5xA1pa. ~8!

Figure 7 showsK520 synthetic ‘‘average’’ echo envelopes calculated using the best-fit model for the silt substdata shown in Fig. 1. These simulations were created uthe average echo of Fig. 1 and the covariance matrixplayed in Fig. 2. The amplitude deviations and degreescorrelation between neighboring samples are realistic,comparison with Fig. 1 confirms. The 20 (Mf ,w2 ,sv) solu-tions yield the following statistics:

Parameter Original Mean Stdv

Parameter

PairCorrelCoeff

Mf 4.68 4.67 0.10 (Mf ,w2) 20.47w2 (cm4! 0.000 91 0.000 92 0.000 22 (Mf ,sv) 20.23

sv (m21! 0.086 0.078 0.003 (w2 ,sv) 20.14

where ‘‘Original’’ refers to the original solution vector. Inthis example the mean values of the Monte Carlo solutiare similar to the original parameters, the standard deviatare a small percentage of the mean values~with possibleexception ofw2), and absolute values of the correlation cefficients are less than 0.5.

In the following sections, plots ofMf andw2 are usedfor distinguishing bottom types. Assuming that the solutioare jointly Gaussian distributed, the 90% error ellipse of (w2

vs Mf) is calculated and plotted in Fig. 8. For this exampthe observed echo variability may account for solution intvals: 4.43<Mf<4.89 and 0.0004<w2<0.0014.

Da--a-

FIG. 7. Monte Carlo simulations for silt substrate in Fig. 1: Model paraeters: f a533 kHz, Mf54.68, g53.3, w250.0009 cm4, sv50.086 m21,uT58°, b51.28.

D. D. Sternlicht and C. P. de Moustier: Remote sensing of sediment

Page 7: Remote sensing of sediment characteristics by optimized echo ...

vedtig

atzeo

mnsoth

ho-fa

iodeHe

co

t

thp

uiteanoth

tic

inte

ephth

ad

heite

dtc

ngo

nu

asty,tomtlews

ates,theffi-

attal

wase.

ti-to

erei-

d at12om-cer

sigh

y isnt

-u-

ntial40-lcru-nddingring

;

IV. SHALLOW-WATER SURVEYS

The sonar system described in Refs. 21, 27 was deoped to evaluate the accuracy of the temporal model anpotential for bottom classification over a range of acousfrequencies and transducer orientations. Circular pistonometries were chosen for their symmetrical directivity pterns, with beamwidths of 21° at 33 kHz and 10° at 93 kHso that for each transducer orientation, an adequate rangbottom incident angles could be insonified by a single shpulse of 0.45 msec at 33 kHz and 0.16 msec at 93 kHz.

For meaningful comparison of model and data, the teporal model utilizes a digitized representation of the tramitted signal, and measured voltage waveforms are cverted to their respective pressure waveforms usingtransducer’s mechanical–electrical transfer function.21,27

A. Survey site

In January and May of 1997, the dual-frequency ecsounder was installed in the instrument well of the 40research vessel ECOS, operated by the Space and NWarfare Systems Center~SPAWAR!. To validate the tempo-ral model and determine the optimum survey configuratfor substrate identification, bottom echoes were recorfrom a range of sediment types with the 33- and 93-ktransducers inclined 0° to 16° from nadir in the roll planData were measured over three sites in San Diego Baysisting, respectively, of sand, silt, and clay substrates.

Bottom characterization was based on:~1! video cover-age recorded during the survey;~2! consulting a sedimendata base for the surrounding area; and~3! analysis of par-ticle size distribution for sediment grabs taken duringsurvey. Sediment samples were separated into size comnents using sieve separation and pipette settling procedoutlined in Ref. 28. The particle size analyses of these sare catalogued in the Appendix, Table IV. At these sites, sparticles constituted the largest grain size percentage; hever, labels of sand, silt, and clay were determined usingcalculatedMf values and observed physical characterisof the samples.

The sand site consisted of a 50-m N–S trackline runnalong the jetty at the mouth of San Diego Bay, in wadepths of 13–15 m, with mean grain size distributionsMf

52, or medium-finesand according to the labeling schemset forth in Ref. 17. The video images revealed an isotrobottom characterized by hillocks with crest–trough heigof 40 cm or more over wavelengths of about 8 m, a ligsprinkling of shell hash, and an occasional starfish or blof kelp.

The silt site consisted of a 150-m N–S trackline of tSan Diego Bay trough—the deepest part of the bay wwater depths of 15–20 m—whose substrate ranged betwclayey sandand sandy mud.17 The video images revealelong stretches of homogeneous substrate, occasional paof kelp, and sole blades of sea grass.

The clay site consisted of a 50-m E–W trackline runnijust north of San Diego Bay’s North Island, water depths11–13 m, with mean grain size distributionsMf57.0, orsandy clay.17 The grain-size analysis identifies this sedimeas borderline silt–clay, but we categorize it as clay beca

J. Acoust. Soc. Am., Vol. 114, No. 5, November 2003 D. D.

l-itsce--,of

rt

--n-e

-tval

nd

z.n-

eo-

ressd

w-e

s

gr

icste

hen

hes

f

tse

of the relatively smooth seascape observed, and the pwaterlogged character of the physical samples. The botvideo revealed a featureless, isotropic bottom, with litflora or fauna except for what appeared to be small burroless than a centimeter in diameter.

The uncomplicated appearance of these three substrthe high spatial overlap between consecutive pings, andgenerally level bathymetry, were conditions deemed sucient for testing the accuracy of the temporal model.

B. Data

The acoustic survey for each site was carried outspeeds of 1–2 kn, ping repetition rate of 5 Hz, and horizondisplacements of about 0.1 m per ping. The transducerelevated to a specified angle from nadir in the roll planAngles of pitch and roll were digitized for each ping repetion and used, along with knowledge of local bathymetry,determine the angle of incidence (uT) of the transducer’smaximum response axis on the bottom. Sea conditions wgenerally mild, with pitch and roll standard deviations typcally less than 0.5°.

Echoes from the San Diego Bay substrates measure33 and 93 kHz are plotted in Figs. 9 and 10. A total ofscenarios is analyzed, each characterized by a unique cbination of acoustic frequency, sediment type, and transduorientation~Appendix, Table V!. It is tempting to interpretthe raster images~Figs. 9, 10! as true geophysical crossections of the bottom; however, penetration at these hacoustic frequencies is limited and the observed energdue primarily to scattering from the water–sedimeinterface.

The raster image of Fig. 9~c! shows a 30-m track segment with a gradual downward slope of the bottom, modlated by the vessel’s heave—whose removal is essefor echo alignment and averaging. In contrast, thecmdepth fluctuations apparent in Fig. 10~a! represent actuatopography. Therefore, these data sets require a level of stiny to identify artifacts that can unfairly bias the shapes aamplitudes of the backscattered echoes. Objects protrufrom or suspended over the bottom may cause scatte

FIG. 8. Scatter plot (w2 vs Mf) and 90% error ellipse for Monte Carlosimulations.~•! Monte Carlo solutions;~n! mean of Monte Carlo solutions~1! original solution.

2733Sternlicht and C. P. de Moustier: Remote sensing of sediment

Page 8: Remote sensing of sediment characteristics by optimized echo ...

r.

FIG. 9. Waterfall and raster plots fo300 consecutive pings of 33-kHz dataLeft: normal incidence, Right: obliqueincidence.y5time in ms since trans-mit, x5ping number.

r

he30toauge

v

ibedhe

choci-

um

anomalies and/or reduced signal levels. The early retuevident in pings 230–280 of Fig. 9~c! are most likely causedby a school of fish swimming near the bottom. Similarly, tscattered energy preceding the bottom profile in pings 4460 of Fig. 10~e! is a strong indication of flora anchoredthe sediment. Bubbles on the face of the transducer can ctemporary dropout of signal amplitude, as evident in pin80–100 of Fig. 10~b!. Data segments clearly exhibiting thartifacts described above are rejected.

Segments of these data sets are combined into an a

2734 J. Acoust. Soc. Am., Vol. 114, No. 5, November 2003

ns

ses

er-

age echo envelope~average pressure vs time! for comparisonwith the temporal model.

V. OPTIMUM FITS OF MODEL WITH DATA

The two-stage parameter estimation technique descrin Sec. III was applied to average echo envelopes from t

12 scenarios presented in Sec. IV B. A group delay ealignment technique was applied to 93-kHz oblique indence measurements made over sand and silt, and minim

D. D. Sternlicht and C. P. de Moustier: Remote sensing of sediment

Page 9: Remote sensing of sediment characteristics by optimized echo ...

r.

FIG. 10. Waterfall and raster plots fo300 consecutive pings of 93-kHz dataLeft: normal incidence, Right: obliqueincidence.y5time in ms since trans-mit, x5ping number.

ta

rbuhea—fr

ataeenpre-the

ca-ure-er

threshold alignment was used for all other scenarios. Deof the echo alignment techniques are given in Ref. 13.

The volume scatter penalty@Eq. ~7!# is applied in all thescenarios, except for 93-kHz oblique incidence measuments on clay and silt. For the latter, large volume contritions are expected to dominate the signal amplitude wtransducer elevation angles are large relative to the bewidth, and in conditions of increased bottom penetrationsuch as water-saturated sediment and/or low acousticquencies.

J. Acoust. Soc. Am., Vol. 114, No. 5, November 2003 D. D.

ils

e--n

m-

e-

For each scenario, approximately ten model–dmatches were determined with 50% or less overlap betwdata segments. A summary of model–data matches issented in Table I. First- and second-order statistics ofresults are listed in the Appendix, Tables VI and VII.

To determine the best prospects for sediment classifition, we evaluated parameter estimates for the four measment combinations~two acoustic frequencies, two transducorientations! and concluded that scatter plots ofw2 vs Mf

effectively delineate the bottom substrates~Figs. 11, 12!.

2735Sternlicht and C. P. de Moustier: Remote sensing of sediment

Page 10: Remote sensing of sediment characteristics by optimized echo ...

value.

2736 J. Acoust. S

TABLE I. Sediment classification summary. Mean values are rounded off to the nearest one-tenthGeoacoustic parameters (r,n,kp) are calculated fromMf with relationships described in Ref. 13.

Site Freq~kHz!Transducerorientation Mf g w2 (cm4) n r

~dB/m/kHz!kp sv (m21)

Sand 33 Normal 3.9 3.00 0.0050 1.041 1.232 0.680 0.20Oblique 3.2 3.00 0.0058 1.072 1.313 0.591 0.04

93 Normal 2.5 3.00 0.0039 1.109 1.458 0.516 0.95Oblique 2.5 3.00 0.0024 1.110 1.464 0.516 0.10

Silt 33 Normal 5.0 3.30 0.0008 1.001 1.170 0.473 0.07Oblique 4.7 3.30 0.0012 1.012 1.186 0.653 0.09

93 Normal 5.2 3.30 0.0007 0.993 1.156 0.363 0.26Oblique 5.3 3.30 0.0022 0.989 1.150 0.308 0.31

Clay 33 Normal 5.3 3.30 0.0005 0.990 1.151 0.329 0.05Oblique 5.3 3.30 0.0006 0.989 1.149 0.316 0.05

93 Normal 5.2 3.30 0.0004 0.992 1.154 0.350 0.10Oblique 5.2 3.30 0.0005 0.993 1.157 0.365 0.17

peheu,dri

de-

ude,e-

arp(

e

rionan-

ard

gh-e.001fore

thes ar

ntaure-

nd

ringsses-akaldi-

ce

A. Mean grain size „Mf… vs relief spectrum strength„w 2…

On the whole, estimated values of mean grain size (Mf)agree with ground-truth measurements presented in Apdix Table IV, and model–data matches for silt exhibit tmost consistency across acoustic frequency and transdorientation. The 33-kHz (Mf) estimates for sand are highapproaching the range characteristic of silts. This may beto local deviation of impedance contrast from the genevalues employed by the model–data matching technique—

FIG. 11. Scatter plot of model–data matches at 33 kHz: Site locations:~s!Sand; ~1! silt; ~* ! clay; ~L! mean value and center of 90% confidenregion ~solid line!. Transducer orientation:~a! Normal; ~b! Oblique.

oc. Am., Vol. 114, No. 5, November 2003

n-

cer

uecan

inference supported by sediment density measurementsscribed in Refs. 29, 30. As the estimated (Mf) parameter isespecially sensitive to changes in measured echo ampliterrors in field calibration may also contribute to disagrements between model–data matches and ground truth.

Mean grain size (Mf) estimates for the clay site~;5.3!are lower than the ground-truth values~.6.5! because thevolume signal component is overestimated, due to the shdecrease of the sediment acoustic attenuation constantkp)correlated with high values ofMf .17 For these fine-grainsediments, accurateMf matching is limited, and applyinglocally determined (r,n,kp) trends would probably producmore realistic model–data matches.

Mean grain size (Mf) estimates for sand exhibit greatevariability than for fines, with measures of standard deviatranging from 0.14 to 0.44, and 90%-confidence regions spning as many as three gradations. In general,Mf estimatesfor silt and clay exhibit more modest ranges, with standdeviations spanning 0.08 to 0.38.

As seen in Table VI and Figs. 11, 12, estimates of rouness spectral strength (w2) are greater than 0.001 for thsand site, less than 0.001 for the clay site, and about 0for the silt site. This trend follows the logic that the relieenergy density spectra of coarse-grain sediments have menergy than those of fine-grain sediments. Variation inestimate ofw2 appears greater in fines than in sands. Apercentage of the mean value,w2 standard deviations fofine-grain sediments~24%–56%! are typically larger thanthose for sand~19%–38%!.

Note from Table VII and Figs. 11, 12 that anticorrelatioof Mf and w2 is also a general bias of the model–damatching procedure. This is especially true of sand measments, where (Mf ,w2) correlation coefficients range from20.58 to 20.96, causing the pronounced slope in the saconfidence regions.

In the literature there is agreement that bottom scattemeasurements can be matched to general bottom cla~fines, sand, gravel, rock!;31 however, correlation of scattering strength to grain size distribution is thought to be wewithin each sediment class. The variability in the individuecho amplitudes that we measured confirms this. If, as in

D. D. Sternlicht and C. P. de Moustier: Remote sensing of sediment

Page 11: Remote sensing of sediment characteristics by optimized echo ...

ao

ba

eals ofmlique

r

thatomgth,

anea-

ciesay

than

sti-incanate

earus-

of

ari-cesd in

-

ing

ce

cated in this analysis, interface roughness characteristicsgrain size distributions were complementary, evaluationecho shape may allow a degree ofintraclassseparation.

B. Sediment volume scattering coefficient „sv…

Estimates of the volume scattering coefficient (sv) areperhaps the most difficult to interpret and, as seen in TaVI, standard deviations on the order of 3 dB from the me

FIG. 12. Scatter plot of model–data matches at 93 kHz: Site locations:~s!Sand; ~1! silt; ~* ! clay; ~L! mean value and center of 90% confidenregion ~solid line!. Transducer orientation:~a! Normal; ~b! Oblique.

J. Acoust. Soc. Am., Vol. 114, No. 5, November 2003 D. D.

ndf

len

value are not uncommon. This variation may be due to rchanges in the statistics governing neighboring patcheseafloor. The roughsv frequency dependencies inferred froour acoustic backscatter measurements at normal and obincidences over sediments in San Diego Bay aref 1.2 for sandand silt, andf 1.0 for clay. These values are slightly highethan, but not inconsistent with thef 0.7 trend inferred from thebackscatter measurements analyzed in Refs. 14, 32. Notean f 4 dependence would indicate Rayleigh scattering frinhomogeneities much smaller than an acoustic wavelenwhereas frequency independence ofsv would imply geomet-ric scattering from inhomogeneities significantly larger ththe acoustic wavelength, as might be the case for our msurements over clay. Our inferred frequency dependenindicate that the volume scatterers in the San Diego Bsediments have a range of sizes both smaller and largerthe acoustic wavelength.

At a given frequency and transducer orientation, emates of sv are reasonably consistent for the fine-grasediments. However, for sand, normal incidence valuesexceed oblique incidence values by 10 dB. This may indica shortcoming in the model assumption thatsv is uniformnear the sediment–water interface. Ifsv increases withdepth, estimated values at normal incidence will applarger than those for oblique incidence. This is due to acotic penetration at normal incidence to depths wheresv islarger—an interpretation consistent with the observationsRefs. 30, 33.

VI. EFFECTS OF ECHO VARIABILITY

Changes in bottom characteristics as well as echo vability due to random constructive/destructive interferenand scattering centers contribute to the observed spreaparameter estimates~Figs. 11, 12!. The length scale of thesurvey and the averaging of 100 pings~corresponding toroughly the along-track extent of the beam’s26-dB foot-print! removes some of the ‘‘natural’’ variability in the individual ping echoes. To investigate the effects ofresidualecho variability on the outputs of the model–data match

13344

3344

3344

TABLE II. Monte Carlo statistics.

Site Freq~kHz!Transducerorientation

OriginalMf

MeanMf

StdvMf

Originalw2

~cm4!

Meanw2

~cm4!

Stdvw2

~cm4!

Originalsv

~m21!

Meansv

~m21!

Stdvsv

~m21! Fig.

Sand 33 Normal 3.72 3.57 0.38 0.005 75 0.006 40 0.001 71 0.201 0.199 0.017~a!

Oblique 3.17 3.21 0.08 0.005 49 0.005 36 0.000 61 0.035 0.037 0.004 1~b!

93 Normal 2.60 2.51 0.28 0.003 49 0.003 69 0.000 62 0.558 0.572 0.150 1~a!

Oblique 2.30 2.36 0.12 0.002 46 0.002 36 0.000 19 0.010 0.045 0.061 1~b!

Silt 33 Normal 4.99 4.99 0.05 0.000 80 0.000 80 0.000 22 0.065 0.069 0.006 1~c!

Oblique 4.68 4.67 0.10 0.000 91 0.000 92 0.000 22 0.086 0.078 0.003 1~d!

93 Normal 5.12 5.11 0.08 0.000 74 0.000 76 0.000 12 0.288 0.305 0.028 1~c!

Oblique 5.23 5.16 0.07 0.001 16 0.000 89 0.000 34 0.226 0.273 0.013 1~d!

Clay 33 Normal 5.27 5.27 0.04 0.000 48 0.000 55 0.000 18 0.059 0.067 0.005 1~e!

Oblique 5.24 5.24 0.05 0.000 66 0.000 78 0.000 36 0.047 0.045 0.003 1~f!93 Normal 5.18 5.20 0.04 0.000 34 0.000 33 0.000 05 0.122 0.122 0.014 1~e!

Oblique 5.19 5.20 0.04 0.000 54 0.000 53 0.000 12 0.181 0.190 0.012 1~f!

2737Sternlicht and C. P. de Moustier: Remote sensing of sediment

Page 12: Remote sensing of sediment characteristics by optimized echo ...

e

er-were

andntee-he

ofthe

lu-ns,

TABLE III. Monte Carlo: Parameter correlation.

Site Freq.~kHz!Transducerorientation (Mf ,w2) (Mf ,sv) (w2 ,sv)

Sand 33 Normal 20.94 10.10 20.19Oblique 20.64 10.55 20.45

93 Normal 20.87 20.32 10.64Oblique 20.85 10.90 20.88

Silt 33 Normal 20.30 20.31 10.02Oblique 20.47 20.44 10.32

93 Normal 20.86 20.28 10.12Oblique 10.42 20.23 20.14

Clay 33 Normal 10.05 20.75 10.04Oblique 10.44 10.07 20.05

93 Normal 20.62 20.66 10.57Oblique 10.26 20.24 20.29

2738 J. Acoust. Soc. Am., Vol. 114, No. 5, November 2003

procedure, (Mf ,w2 ,sv) solutions close to the mean valufor each~substrate, frequency, orientation! combination werechosen. Then, for each original solution, 20 synthetic avage echo envelopes and matched parameter solutionsgenerated as described in Sec. III C.

For each measurement scenario, an original solutionthe mean and standard deviation of the corresponding MoCarlo solution are summarized in Table II. Correlation btween parameter pairs are summarized in Table III. TMonte Carlo solutions forw2 vs Mf are shown in Figs. 13and 14 for 33 and 93 kHz, respectively. The distributionsMonte Carlo solutions are adequately represented byconfidence regions—with the exception of Fig. 13~b! ~sand,33 kHz, oblique! which demonstrates one-sidedw2 cluster-ing about 0.0055. The mean values of the Monte Carlo sotions are in general agreement with the original solutiowith the exception of Fig. 14~d! ~silt, 93 kHz, oblique!.

s

FIG. 13. Scatter plot of Monte Carlosolutions at 33 kHz: Panel descriptionin Table II. ~•! Monte Carlo solutions;~n! mean value and center of 90%confidence region;~s! original sandsolution;~1! original silt solution;~* !original clay solution.

D. D. Sternlicht and C. P. de Moustier: Remote sensing of sediment

Page 13: Remote sensing of sediment characteristics by optimized echo ...

s

FIG. 14. Scatter plot of Monte Carlosolutions at 93 kHz: Panel descriptionin Table II. ~•! Monte Carlo solutions;~n! mean value and center of 90%confidence region;~s! original sandsolution;~1! original silt solution;~* !original clay solution.

heth

-faei

be

aar

esromtanr

d at

thengicignalare

entsty

po-

eene

allythe

In Figs. 15 and 16, 90%-confidence regions for tMonte Carlo solutions are juxtaposed with those formodel–data solutions reported in Sec. V~henceforth calledthe ‘‘real’’ solutions! at 33 and 93 kHz, respectively. In general, the confidence regions of the Monte Carlo solutionswithin those of the real solutions. However, the plots suggthat for normal incidence over sand at 33 kHz, variationsMf andw2 are larger than suggested by the limited numof field measurements. The same can be said ofw2 estimatesfor oblique incidence over clay at 33 kHz. This implies thanalysis of larger data sets could yield greater solution vability than what is currently observed.

As observed for the real solutions, the Monte Carlotimates ofMf for sand exhibit greater variability than fofines—with measures of standard deviation ranging fr0.08 to 0.38 and 0.04 to 0.10, respectively. As a percenof the mean value,w2 standard deviations for fine-graisimulations~15%–46%! are typically larger than those fo

J. Acoust. Soc. Am., Vol. 114, No. 5, November 2003 D. D.

e

llstnr

ti-

-

ge

sand ~8%–27%!. Standard deviations forsv are typicallyless than 10% of the mean value, with exceptions for san33 kHz.

Also, a significant anticorrelation betweenMf and w2

for sand substrates is seen in the real solutions and inMonte Carlo solutions, with correlation coefficients rangifrom 20.64 to 20.94. In the temporal model of acoustbackscatter, increasing either parameter decreases speak amplitude, and vice versa. In nature, these quantitiesexpected to be negatively correlated—i.e., coarser sedim~lower Mf) exhibit more energy in the relief energy densispectrum~larger w2). When the ‘‘true’’ signal is contami-nated by ‘‘noise’’ the parameters also tend to adjust in opsite directions.

There also appears to be modest anticorrelation betwMf and sv in solutions for fine-grain sediments. In thessubstrates, scattering from the sediment volume typicplays a larger role. An increase in either parameter raises

2739Sternlicht and C. P. de Moustier: Remote sensing of sediment

Page 14: Remote sensing of sediment characteristics by optimized echo ...

twttlr-

geet

ssenor

tist-

eliz

utll

ri-u

nain

erea

a-

ent

am-by a

pe

ean

o, the

de,re-glo-u-thegth,nedrre-oesromionandruth

omd as-sess-the

iveribeditesam-andrac-

s ofina-tion

rac-h

-leen-ons-resredal–

i-n of

;

calculated energy in the signal tail, and apparently theparameters compete to fit this section of the signal. Too liis known aboutin situ sediment volume scattering characteistics to warrant a physical interpretation.

In theory, values of the data covariance matrix~and thusthe solution variance! can be decreased by averaging a larnumber of echoes. However, in our data sets, processingsembles much greater than 100 pings excessively filtersshape characteristics of the average envelope that are etial to the matching procedure. Furthermore, with largesembles the requirement of bottom homogeneity is mlikely to be violated—especially at high survey speeds.

In theory also, the data covariance matrix has potenapplication in model–data fitting. The nonweighted leasquares merit function of Eq.~5! was chosen over a varianceweighted approach in order to favor peak amplitude moddata matching—emphasizing extraction of mean grain scorrelated parameters, such as impedance contrast. Fwork with this technique will include testing of the fumaximum likelihood estimation~MLE! paradigm; i.e., cova-riance matrix weighting of the model–data disparity. Vaance weighting of each model–data sample disparity shoimprove model–data fitting at the leading and trailing sigedges—at the expense of precise peak amplitude matchThe effect, however, of the data’s covariances should cothe optimized model to assume the true ‘‘shape’’ of the m

FIG. 15. Confidence regions for Monte Carlo and real solutions~33 kHz!.Real solutions: Solid lines ———90%-confidence regions;~L! mean val-ues. Monte Carlo solutions: Dashed lines 90%-confidence regions~n! mean values;~s! original sand solution;~1! original silt solution;~* !original clay solution. Transducer orientation:~a! Normal; ~b! Oblique.

2740 J. Acoust. Soc. Am., Vol. 114, No. 5, November 2003

oe

rn-

heen--e

al-

–eure

ldlg.

ce-

sured envelope. Comparisons between full MLE optimiztion and peak amplitude~nonweighted! matching will beevaluated in future evolutions of this echo envelope sedimcharacterization algorithm.

VII. SUMMARY

The method for estimating sediment geoacoustic pareters presented here compares bottom returns measuredcalibrated, moderate beamwidth~10°–21°!, vertically ori-ented ~0°–15°! monostatic sonar, with an echo envelomodel based on high-frequency~10–100 kHz! incoherentbackscatter theory and sediment properties such as mgrain size (Mf), interface roughness (w2 ,g), and sedimentvolume scattering statistics (sv). A two-stage average echenvelope matching procedure was described where: firstsediment type~sand or fines! is established by iterating onthe reflection coefficient to match the peak echo amplituand to establish a general fit with generic values of themaining geoacoustic parameters; then, a three-parameterbal optimization is performed using a combination of simlated annealing and downhill simplex searches overallowable range of interface roughness spectral strensediment volume scattering coefficient, and a constrairange of reflection and bottom absorption coefficients colated to mean grain size. In San Diego Bay, bottom echwere collected at 33 and 93 kHz over substrates ranging fsand to clay. Application of the sediment characterizatmethod to these data yielded solutions for grain sizegeoacoustic properties that are consistent with ground-tmeasurements.

The ground-truth measurements, consisting of bottvideo, grain size analyses, environmental databases, ansociated ranges of geoacoustic parameters, lack direct asments of the modeled geoacoustic properties. This, andsmall number of sites and regions evaluated, limits definitassessment of the accuracy and robustness of the descinversion technique. Controlled, calibrated surveys over scharacterized for the complete range of geoacoustic pareters must eventually be employed to further evaluateimprove the efficacy of the echo envelope sediment chaterization technique.

For the experiments described in this paper, analysethe estimated geoacoustic parameters for different combtions of sediment type, frequency, and transducer orientasuggest that moderate frequencies~33 kHz! and normal inci-dence are more suitable for this method of sediment chaterization. This may, in part, be due to limitations at higacoustic frequencies~e.g., 93 kHz! of the backscatter model’s underlying Kirchhoff theory, and partly due to the simptemporal structure of the returns at lower acoustic frequcies ~33 kHz!—simplifying calculation of the average echenvelope. Furthermore, approximate alignment of the traducer’s maximum response axis at normal incidence insuthat the maximum interface component of the backscattesignal will exceed the maximum volume contribution—condition necessary for reducing ambiguity in the modedata matching procedure.

The ability to distinguish sands from fine-grain sedments was demonstrated based on acoustic estimatio

D. D. Sternlicht and C. P. de Moustier: Remote sensing of sediment

Page 15: Remote sensing of sediment characteristics by optimized echo ...

ros

c-ofrvee-

fc

am-de-ithis

areilityid-

ticugh-ean-

ga-umin

ree

at isa of

msicn of

fes-iser

by14-

;

mean grain size alone. The creation of feature vectors festimates of mean grain size (Mf) and interface roughnesspectral strength (w2) showed promise forintraclasssepara-tion of silt and clay. Limitations on interface curvature ditated by the Kirchhoff approximation restrict applicationthis scheme to sediments having a large rms radius of cuture relative to the acoustic wavelength. This excludestremely rough~rocky! substrates, or operation at high fr

FIG. 16. Confidence regions for Monte Carlo and real solutions~93 kHz!.Real solutions: Solid lines ———90%-confidence regions;~L! mean val-ues. Monte Carlo solutions: Dashed lines 90%-confidence regions~n! mean values;~s! original sand solution;~1! original silt solution;~* !original clay solution. Transducer orientation:~a! Normal; ~b! Oblique.

APPENDIX: SUPPORTING TABLES

J. Acoust. Soc. Am., Vol. 114, No. 5, November 2003 D. D.

m

a-x-

quencies ~.100 kHz!. Furthermore, local deviations osediment:water impedance ratio~rn! and sediment acoustiattenuation constant (kp) from generic values~mean valuescorrelated withMf) will result in estimates ofMf , w2 , andsv that are distorted from their true values.

Monte Carlo simulations based on a geoacoustic pareter solution set and the data’s covariance matrix werescribed. In the mean, the Monte Carlo solutions agree wthe original solution; however, for a given substrate, thereas much variability in the Monte Carlo solutions as therein an ensemble of real solutions. Therefore, echo variabmust be considered during parameter optimization by proving confidence limits on the results.

According to the observed spread of geoacousmatches from measured signals and synthetic data, roness spectral strength estimates (w2) for sand substrates arrelatively immune to raw echo variability, whereas megrain size estimates (Mf) are moderately affected. The opposite is observed for fine-grain substrates:Mf estimates arerelatively immune to raw echo variability, whereasw2 esti-mates are significantly affected. A more thorough investition of echo envelope averaging procedures and maximlikelihood model–data matching techniques may resultmethods to reduce the (Mf ,w2) confidence regions.

Finally, the classification procedure introduces a degof anticorrelation betweenMf and w2 , which is especiallylarge for sand substrates. This trend is consistent with whexpected in nature, where the relief energy density spectrcoarser sediments~lower Mf! exhibit more energy~higherW2! than those of fine-grain substrates.

ACKNOWLEDGMENTS

We are grateful to the Space and Naval Warfare SysteCenter~SPAWAR! for their assistance in collecting acoustdata and bottom samples, Reson Inc. for the generous loasonar transducers, and ORINCON Defense for their prosional support. We thank Professor William Coles for hinput on conducting the error analysis, and Jo Griffith for hhelp in preparing the illustrations. This work was fundedthe Office of Naval Research under Contract No. N00094-1-0121.

f and a

TABLE IV. Survey site ground truth. Substrate percentages may not add exactly to 100 due to round-ofsmall gravel constituency.

SiteSampleindex

Latitudedeg minNorth

Longitudedeg min

WestMean grainsize ~PHI!

Mean grainsize ~mm!

%Sand

%Silt

%Clay

Sand 1 32 40.760 117 13.653 1.9 268 93 2 42 32 40.650 117 13.585 2.2 218 93 2 43 32 40.647 117 13.626 1.7 308 90 3 4

Silt 1 32 42.265 117 13.927 4.1 58 76 14 102 32 41.887 117 14.153 5.9 17 49 30 21

Clay 1 32 42.997 117 11.728 6.5 11 39 33 272 32 42.995 117 11.767 6.8 9 36 34 293 32 42.997 117 11.814 6.6 10 38 33 29

2741Sternlicht and C. P. de Moustier: Remote sensing of sediment

Page 16: Remote sensing of sediment characteristics by optimized echo ...

ich the

-

22032

0607

2503

2742 J. Acoust. S

TABLE V. Survey measurement characteristics. Normal incidence measurements refer to those for whcentral axis of the transducer’s radiation pattern is aligned with the bottom normal~meanuT;0). Obliqueincidence measurements exhibit elevation angles roughly corresponding tou3dB/2 ~8°–12° and 6°–9°, respectively, at 33 and 93 kHz!. Columns 3–6 represent mean or approximate values.

Site Freq~kHz!Transducerorientation

Transducerelevation

angleuT (deg)

Transduceraltitudealt (m)

Along-track3-dB

footprintD3 (m)

Along-track6-dB

footprintD6 (m)

Transmissionsource leveldB re: 1 mPa

@ 1 mFigurelabel

Sand 33 Normal 2.0 13.5 5.0 7.0 197.8 9~a!Oblique 12.0 13.5 5.0 7.0 197.8 9~b!

93 Normal 2.0 13.5 2.4 3.1 191.4 10~a!Oblique 6.5 14 2.5 3.2 191.4 10~b!

Silt 33 Normal 2.0 16 5.9 8.3 192.4 9~c!Oblique 8.0 16.5 6.1 8.5 192.4 9~d!

93 Normal 2.0 19 3.3 4.3 192.3 10~c!Oblique 8.5 19 3.3 4.3 192.3 10~d!

Clay 33 Normal 0 12 4.5 6.2 197.8 9~e!Oblique 12.0 11 4.1 5.7 197.8 9~f!

93 Normal 1.0 12.5 2.2 2.8 191.4 10~e!Oblique 7.0 13 2.3 3.0 191.4 10~f!

TABLE VI. Sediment classification statistics.

Site Freq~kHz!Transducerorientation

MeanMf

StdvMf g

Meanw2

~cm4!Stdv w2

~cm4!Meansv

~m21!Stdv sv

~m21!

Sand 33 Normal 3.88 0.28 3.00 0.004 96 0.001 13 0.202 0.0Oblique 3.16 0.14 3.00 0.005 77 0.001 09 0.041 0.02

93 Normal 2.48 0.44 3.00 0.003 91 0.001 10 0.945 0.34Oblique 2.45 0.27 3.00 0.002 36 0.000 90 0.104 0.11

Silt 33 Normal 4.96 0.09 3.30 0.000 76 0.000 20 0.071 0.01Oblique 4.65 0.22 3.30 0.001 15 0.000 32 0.091 0.01

93 Normal 5.18 0.09 3.30 0.000 72 0.000 19 0.261 0.07Oblique 5.32 0.38 3.30 0.002 23 0.001 24 0.314 0.22

Clay 33 Normal 5.26 0.08 3.30 0.000 48 0.000 21 0.048 0.02Oblique 5.30 0.20 3.30 0.000 61 0.000 24 0.047 0.01

93 Normal 5.21 0.15 3.30 0.000 41 0.000 11 0.102 0.04Oblique 5.18 0.24 3.30 0.000 55 0.000 13 0.170 0.05

ering

for

e

ticrties6,

r forUn-

for16thical9–

o-

TABLE VII. Sediment classification: Parameter correlation.

Site Freq.~kHz!Transducerorientation (Mf ,w2) (Mf ,sv) (w2 ,sv)

Sand 33 Normal 20.86 20.79 10.37Oblique 20.72 20.78 20.48

93 Normal 20.96 20.79 10.77Oblique 20.58 10.66 10.07

Silt 33 Normal 10.19 10.07 20.6Oblique 20.27 10.17 20.74

93 Normal 20.23 20.87 10.44Oblique 20.58 20.07 10.60

Clay 33 Normal 10.18 20.50 20.11Oblique 20.31 20.64 20.2

93 Normal 20.34 20.77 10.17Oblique 10.77 10.26 10.07

oc. Am., Vol. 114, No. 5, November 2003

1N. G. Pace and R. V. Ceen, ‘‘Seabed classification using the backscattof normally incident broadband acoustic pulses,’’ Hydrographic J.26,9–16 ~1982!.

2R. Chivers, N. Emerson, and D. R. Burns, ‘‘New acoustic processingunderway surveying,’’ Hydrographic J.56, 9–17~1990!.

3L. A. Mayer. UNB-OMG/UNH-CCOM Multibeam Sonar Traning CoursNotes, November 2002.

4A. S. Tsehmahman, W. T. Collins, and B. T. Prager, ‘‘Acousseabed classification and correlation analysis of sediment propeby QTC view,’’ in Proceedings of IEEE OCEANS 97, pp. 921–921997.

5G. J. Heald and N. G. Pace, ‘‘An analysis of 1st and 2nd backscatteseabed classification,’’ in Proceedings of III European Conference onderwater Acoustics, Crete, June 1996, pp. 649–654.

6G. J. Heald and N. G. Pace, ‘‘Implications of bi-static treatmentthe second echo from a normal incidence sonar,’’ in Proceedings:International Congress on Acoustics and 135th Meeting AcoustSociety of America, Seattle, Washington, June 1998, Vol. IV, pp. 3003010.

7S. G. Schock, L. R. LeBlanc, and L. A. Mayer, ‘‘Chirp subbottom prfiler for quantitative sediment analysis,’’ Geophysics54~4!, 445–450~1989!.

D. D. Sternlicht and C. P. de Moustier: Remote sensing of sediment

Page 17: Remote sensing of sediment characteristics by optimized echo ...

t

m

acng

ca

de

-

oo

he,’’

ts,

cy

ntaS

ts,

oftio

to-a

h

erp.ch

y

er,m.

ver

andch.

a-o,

di-.S.pi,

delsEE

nan.

h-

tom

ng:Am.

8L. R. LeBlanc, L. Mayer, M. Rufino, and J. King, ‘‘Marine sedimenclassification using the chirp sonar,’’ J. Acoust. Soc. Am.91~1!, 107–115~1991!.

9M. V. Berry, ‘‘The statistical properties of echoes diffracted frorough surfaces,’’ Philos. Trans. R. Soc. London, Ser. A273, 611–654~1973!.

10E. H. Nesbitt, ‘‘Estimation of sea bottom parameters using acoustic bscattering at vertical incidence,’’ Master’s thesis, University of Washiton, 1988.

11D. R. Jackson and E. Nesbitt, ‘‘Bottom classification using backstering at vertical incidence,’’ J. Acoust. Soc. Am. Suppl. 183, S80~1988!.

12X. Lurton and E. Pouliquen, ‘‘Identification de la nature du fondla mer al’aide de signaux d’e´cho-sondeurs. II. Me´thode d’identificationet resulatats experimentaux,’’ Acta Acust.~European Acoustics Association! 2~3!, 187–194~1994!.

13D. D. Sternlicht and C. P. de Moustier, ‘‘Time-dependent seaflacoustic backscatter~10–100 kHz!,’’ J. Acoust. Soc. Am.114, 2709–2725~2003!.

14D. R. Jackson, D. P. Winebrenner, and A. Ishimaru, ‘‘Application of tcomposite roughness model to high-frequency bottom backscatteringAcoust. Soc. Am.79~5!, 1410–1422~1986!.

15C. K. Wentworth, ‘‘A scale of grade and class terms for clastic sedimenJ. Geol.30~5!, 377–392~1922!.

16W. C. Krumbein, ‘‘Application of logarithmic moments to size frequendistribution of sediments,’’ J. Sediment. Petrol.6, 35–47~1936!.

17Applied Physics Laboratory. High-Frequency Ocean EnvironmeAcoustic Models Handbook. Technical report, APL-UW TR9407, AEA9501 University of Washington, 1994.

18E. L. Hamilton, ‘‘Compressional wave attenuation in marine sedimenGeophysics37, 620–646~1972!.

19Z. H. Michalopoulou, D. Alexandrou, and C. de Moustier, ‘‘Applicationa maximum likelihood processor to acoustic backscatter for the estimaof seafloor roughness parameters,’’ J. Acoust. Soc. Am.95~5!, 2467–2477~1994!.

20H. Matsumoto, R. P. Dziak, and C. Fox, ‘‘Estimation of seafloor micropographic roughness through modeling of acoustic backscatter datcorded by multi-beam systems,’’ J. Acoust. Soc. Am.94~5!, 2776–2787~1993!.

21D. D. Sternlicht and C. P. de Moustier, ‘‘Temporal modeling of hig

J. Acoust. Soc. Am., Vol. 114, No. 5, November 2003 D. D.

k--

t-

r

J.

’’

l

’’

n

re-

frequency ~30–100 kHz! acoustic seafloor backscatter: Shallow watresults,’’ in High Frequency Acoustics in Shallow Water, CP-45, p509–516, Lerici, Italy, July 1997. NATO SACLANT Undersea ResearCentre.

22G. E. Forsythe, M. A. Moler, and B. Cleve,Computer Methods forMathematical Computations~Prentice-Hall, Englewood Cliffs, NJ,1977!.

23S. Kirkpatrick, C. D. Gelatt, Jr., and M. P. Vecchi, ‘‘Optimization bsimulated annealing,’’ Science220~4598!, 671–680~1983!.

24M. Metropolis, A. Rosenbluth, M. Rosenbluth, A. Teller, and E. E. Tell‘‘Equations of state calculations by fast computing machines,’’ J. ChePhys.21, 1087–1091~1953!.

25W. H. Press and S. A. Teukolsky, ‘‘Simulated annealing optimization ocontinuous spaces,’’ Comput. Phys.5, 426–429~1991!.

26S. Geman and D. Geman, ‘‘Stochastic relaxation, Gibbs distributions,the Bayesian restoration of images,’’ IEEE Trans. Pattern Anal. MaIntell. 6, 721–741~1984!.

27D. D. Sternlicht, ‘‘High Frequency Acoustic Remote Sensing of Sefloor Characteristics,’’ Ph.D. thesis, Univeristy of California, San Dieg1999.

28R. H. Plumb, ‘‘Procedure for handling and chemical analysis of sement and water samples,’’ Technical Report TR-EPA/CE-81-1, UArmy Engineer Waterways Experiment Station, Vicksburg, Mississip1981.

29P. D. Mourad and D. R. Jackson, ‘‘High frequency sonar equation mofor bottom backscattering and forward loss,’’ in Proceedings of IEOCEANS 89, pp. 1168–1175, 1989.

30A. P. Lyons and T. H. Orsi, ‘‘The effect of a layer of varying density ohigh-frequency reflection, forward loss, and backscatter,’’ IEEE J. OceEng.23~4!, 411–422~1998!.

31S. Stanic, K. B. Briggs, P. Fleischer, W. B. Sawyer, and R. I. Ray, ‘‘Higfrequency acoustic backscattering,’’ J. Acoust. Soc. Am.85~1!, 125–136~1989!.

32D. R. Jackson, J. J. Crisp, and P. A. Thompson, ‘‘High-frequency botbackscatter measurements in shallow water,’’ J. Acoust. Soc. Am.80~4!,1188–1199~1986!.

33D. R. Jackson and K. B. Briggs, ‘‘High-frequency bottom backscatteriRoughness versus sediment volume scattering,’’ J. Acoust. Soc.92~2!, 962–977~1992!.

2743Sternlicht and C. P. de Moustier: Remote sensing of sediment