18.11.2014 Views

AIDJEX Bulletin #40 - Polar Science Center - University of Washington

AIDJEX Bulletin #40 - Polar Science Center - University of Washington

AIDJEX Bulletin #40 - Polar Science Center - University of Washington

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

DATA BUOYS<br />

<strong>AIDJEX</strong> BULLETIN No. 40<br />

June 0978<br />

SUMMARY OF TECHNICAL DEVELOPMENTS IN <strong>AIDJEX</strong><br />

--Pat Martin. ........................... 1<br />

ARCTIC ODYSSEY--FIVE<br />

YEARS OF DATA BUOYS IN <strong>AIDJEX</strong><br />

--Pat Martin and C. R. Gillespie .................. 7<br />

ARCTIC ENVIRONMENTAL BUOY SYSTEM<br />

--Walter P. Brown and Edmund G. Kerut ................ 15<br />

AIR DROPPABLE RAMS (ADRAMS) BUOY<br />

--W. P. Brown and E. G. Kerut .................... 21<br />

THE SYNRAMS ICE STATION<br />

--Samuel P. Burke and Beaumont M. Buck ............... 29<br />

PERFORMANCE OF MET/OCEAN BUOYS IN <strong>AIDJEX</strong><br />

--M. G. McPhee, L. Mangum, and P. Martin .............. 35<br />

A TEST OF BAROMETRIC PRESSURE AND TEMPERATURE MEASUREMENTS<br />

FROM ADRAMS BUOYS<br />

--Pat Martin and Me1 Clarke .................... 61<br />

POSITION MEASUREMENTS OF <strong>AIDJEX</strong> MANNED CAMPS<br />

USING THE NAVY NAVIGATION SATELLITE SYSTEM<br />

--Pat Martin, C. G. Gillespie, Alan Thorndike, and D. Wells .... 83<br />

AIRCRAFT LOGISTICS AND DRIFTING BUOY COVERAGE<br />

FOR THE FIRST GARP GLOBAL EXPERIMENT<br />

--Pat Martin. ........................... 103<br />

ET ALII<br />

MATHEMATICAL CHARACTERISTICS OF A PLASTIC MODEL<br />

OF SEA ICE DYNAMICS<br />

--Robert S. Pritchard and R. Reimer ................ 109<br />

A RESEARCH PLAN TO TEST THE <strong>AIDJEX</strong> MODEL<br />

AS AN ICE FORECASTING AID<br />

--Robert S. Pritchard and Max Coon ................. 153<br />

FINAL DISPOSITION OF <strong>AIDJEX</strong> DATA BANK FILES<br />

--Murray Sta teman ......................... 190


<strong>AIDJEX</strong> BULLETIN No. 40<br />

June 1978<br />

Financial support for <strong>AIDJEX</strong> was provided by<br />

the National <strong>Science</strong> Foundatian,<br />

the Office <strong>of</strong> NavaZ Research,<br />

and other U.S. and Canadian agencies.<br />

Arctic Ice Dynamics Joint Experinent<br />

Di vi si on <strong>of</strong> Flari ne Resources<br />

<strong>University</strong> <strong>of</strong> <strong>Washington</strong><br />

Seattle, <strong>Washington</strong> 98105


NOTE TO FRIENDS,<br />

AND BYSTANDERS . . .<br />

PARTICIPANTS, CONTRIBUTORS, COLLABORATORS,<br />

From the beginning <strong>of</strong> the project we felt that, in light<br />

<strong>of</strong> its high cost and risk <strong>of</strong> failure inherent to all enterprises<br />

involving sea ice camps, it was necessary to maintain an<br />

intense level <strong>of</strong> reporting. Since its first issue, published<br />

in 1970, the <strong>AIDJEX</strong> <strong>Bulletin</strong> has been a vehicle for communicating<br />

plans, events , data, analyses, and interpretations. It was<br />

used not only by the participants in the project, but in many<br />

instances also by researchers wishing to communicate their<br />

results rapidly and in a fashion that might benefit the progress<br />

<strong>of</strong> <strong>AIDJEX</strong>.<br />

We have ample correspondence from people saying they will<br />

miss the familiar, rough but vigorous document <strong>of</strong> work in progress,<br />

cheaply bound, with the scenic blue cover. But the project is<br />

done, according to plan, without accident; the data are banked,<br />

and whatever good ideas they proved or inspired will become part<br />

<strong>of</strong> the corporate memory <strong>of</strong> science. The work goes on and there<br />

has been no time for searching our minds for ways we could have<br />

done it better or for congratulating ourselves or, least <strong>of</strong> all,<br />

for waiting to be congratulated.<br />

An old adage <strong>of</strong> science goes: .If you have done some bad or<br />

meaningless research and nobody pays attention, praise your luck<br />

and proceed. .If you have done some flashy but dubious research,<br />

praise your fate, too. You will be talked about and after awhile<br />

people will only remember that you were talked about and forget<br />

that your work was dubious. .If you have done some sound and<br />

fundamental work, it will be so clear and self-evident in people's<br />

minds that, quite naturally, giving credit will become irrelevant.<br />

If you have that kind <strong>of</strong> luck, praise your fate, too, for that<br />

is the highest form <strong>of</strong> accomplishment.<br />

Such reflections aside, it is time to bring the <strong>AIDJEX</strong> <strong>Bulletin</strong><br />

to an end. The only certain thing is that we owe gratitude to<br />

those who have contributed, and to its editor, Alma Johnson. Her<br />

dedication and skills, to a large degree, have made it the lively<br />

document to which we bid farewell with this final issue.<br />

Norbert Untersteiner


Dear Alan, Austin,<br />

Andy, AZexei, Ann,<br />

AZ, Ailsa, Bill,<br />

Bob, Beau, Ben,<br />

C Z ay ton, Clarence,<br />

Charlie, Co Zin,<br />

Dick, Dan, Drew,<br />

Dennis, Dean, Don,<br />

Dave, Erik, Ed,<br />

Eric, Frank, Fred,<br />

Gil lespie, Gary,<br />

George, Gunter,<br />

Gordon, Heinrich,<br />

Hilda, Hans, Igor,<br />

John, Jack, Jim,<br />

Judy, Joe, Jay,<br />

Ken, Konrad, Leo,<br />

Larry, Lon, Mark,<br />

Moira, Max, Miles,<br />

Mort, Mike,<br />

Norbert, Per,<br />

Price, Paul, Ron,<br />

Roger, Rich, Reid,<br />

Rene, Ruby, Sam,<br />

Steve, SeeZye,<br />

Ted, Tony, Willy,<br />

Warren, WaZZy,<br />

and Walt, to name<br />

but a few--<br />

fiat can I say?<br />

The end? No more<br />

where this came from?<br />

Th-th-th-that 's all,<br />

folks? It's over?<br />

It is, you know.<br />

Putting out the <strong>Bulletin</strong> was more fun than anyone said it would be, and<br />

perhaps more fun than it should have been, but I'm glad we did it that way.<br />

Most <strong>of</strong> you I hadn't even met before <strong>AIDJEX</strong> began, and now so many <strong>of</strong> you<br />

are fr-knds. If this last BulZetin means the last assodation we have with<br />

each other, I shall be very sorry. I have enjoyed being your editor.<br />

Best regards<br />

I


SUMMARY OF TECHNICAL DEVELOPMENTS IN <strong>AIDJEX</strong><br />

Pat Martin<br />

<strong>AIDJEX</strong><br />

Early in the planning for <strong>AIDJEX</strong> it was recognized that the success <strong>of</strong><br />

the program depended on the availability <strong>of</strong> reliable automatic data acquisition<br />

and transmission systems (data buoys) and accurate positioning systems.<br />

These systems would involve an extension <strong>of</strong> existing technologies, including<br />

adaptation to arctic operating conditions. Since these development tasks<br />

were <strong>of</strong> central importance to the experiment, and were not within the scope<br />

<strong>of</strong> any single research component, responsibility for them was assigned to<br />

the project <strong>of</strong>fice. As <strong>AIDJEX</strong> Technical Coordinator, I was responsible for<br />

both development efforts. The approach adopted was to contract for the<br />

system's engineering development and production, with the project <strong>of</strong>fice<br />

staff being responsible for stating system requirements, monitoring contractor<br />

performance, and deploying and operating the systems. About ten person-years<br />

<strong>of</strong> effort were devoted to these tasks under the <strong>AIDJEX</strong> <strong>of</strong>fice contract with<br />

the National <strong>Science</strong> Foundation.<br />

Development efforts for the two systems commenced in 1971 for the 1972<br />

pilot study. The primary positioning system chosen for the pilot study was<br />

the Navy Navigation Satellite System, Transit. The necessary equipment was<br />

available commercially and was leased because <strong>of</strong> the short duration <strong>of</strong> the<br />

1972 program. The problems encountered generally resulted from the use <strong>of</strong><br />

relatively new equipment in an unusual environment without sufficient redundancy.<br />

An acoustic positioning system also was acquired in 1972 to resolve<br />

small amplitude (meters), high frequency (hours) ice motion. This system,<br />

while built to the specifications <strong>of</strong> the project, was well within the state<br />

<strong>of</strong> the art and performed as expected. Oscillations <strong>of</strong> ice motion with a<br />

12-hour period observed by Hunkins on T-3 were found to be inertial rather<br />

than tidal with this acoustic positioning system. The other major finding<br />

was that there is little "power" at frequencies <strong>of</strong> ice velocity greater than<br />

2 day-'.<br />

1


Since the 1972 data buoy requirements could not be met with equipment<br />

available commercially, it was necessary to contract for the development and<br />

production.<br />

This effort was carried out in only a few months, and much <strong>of</strong><br />

the final testing was completed on the ice. This pattern was to repeat<br />

itself in the main experiment with unfortunate consequences.<br />

Five <strong>of</strong> six<br />

buoys which were deployed measured and recorded hourly values <strong>of</strong> barometric<br />

pressure successfully for more than a month, providing data to drive calcu-<br />

lations <strong>of</strong> geostrophic wind.<br />

recovery by radio from overflying aircraft.<br />

These buoys were designed to permit data<br />

An additional element <strong>of</strong> the<br />

data buoy program in 1972 was the development and deployment <strong>of</strong> six buoys<br />

which also measured pressure and temperature and communicatedthesedata and<br />

position information through a NASA meteorological satellite, Nimbus 6.<br />

These buoys, with an average life in excess <strong>of</strong> one year, were developed with<br />

NGAY Zudl;ag in response to AIDJFX long-term needs.<br />

The experience with iata<br />

buoys and positioning systems in 1972 was vital to preparations for the main<br />

experiment .<br />

Two <strong>of</strong> the key decisions made in preparing for the 1975-76 main experi-<br />

ment were to contract for Transit systems with additional data logging<br />

capability rather than develop them in-house, and to use two data buoy<br />

tems, one <strong>of</strong> which would not depend on NASA experimental satellites to<br />

transmit data.<br />

The former decision was implemented in the issuance <strong>of</strong> a<br />

request for proposals for NavSat Data Acquisition Systems in late 1973.<br />

technical content <strong>of</strong> the specifications reflected three years <strong>of</strong> intensive<br />

study and experience with Transit.<br />

successfully negotiated by the end <strong>of</strong> 1973.<br />

The vendor was selected and a contract<br />

adaptation <strong>of</strong> commercially available equipment and its assembly into systems<br />

configured to meet the needs <strong>of</strong> the experiment.<br />

<strong>of</strong> the contract the company lost several key people, which caused delays in<br />

design and construction.<br />

postponed about six months until the fall <strong>of</strong> 1974.<br />

sys-<br />

These tests were success-<br />

ful in demonstrating most <strong>of</strong> the system characteristics, including accurate<br />

positioning and data logging.<br />

However , some problems remained, including an<br />

unacceptable temperature sensitivity in the reference oscillators and unfin-<br />

ished programing.<br />

The contract called for the<br />

Shortly after the signing<br />

As a result, field tests <strong>of</strong> the system had to be<br />

The level <strong>of</strong> detail in the contract, together with<br />

substantial amounts <strong>of</strong> money held for progress payments at performance<br />

The<br />

2


milestones, kept the vendor working on these problems,which were not resolved<br />

fully until well into the main experiment.<br />

The performance <strong>of</strong> the NavSat Data Acquisition Systems in the main<br />

experiment was excellent. Position measurements good to about 20 m were<br />

obtained throughout the experiment, the longest interruption being about five<br />

days at one camp. The data logging function also performed well, with occasional<br />

interruptions <strong>of</strong> only a few days. The oceanographic data have some<br />

noise, introduced in the signal processors before the data reached the data<br />

acquisition system, which requires filtering in the data processing. This<br />

noise was present during the field test, but was removed by careful shielding.<br />

Less care was taken in the main experiment. More expensive digitizing hardware<br />

might reduce the noise problem, but the data on hand are adequate.<br />

The acoustic positioning system was used again during the main experiment<br />

to investigate inertial oscillations, and performed as expected. The<br />

total expense for positioning systems during <strong>AIDJEX</strong> was about $600,000.<br />

The early decision to develop data buoys that did not dependonan experimental<br />

satellite not yet launched presented a major technical challenge. The<br />

requirement <strong>of</strong> several position measurements per day, accurate to about 100 my<br />

meant that these would be the first buoys to make extensive use <strong>of</strong> NavSat.<br />

The requirement for data transmission to a central station could only be met<br />

by high frequency radio, which increased the power consumption and buoy complexity.<br />

Against these demands, redundant measurements <strong>of</strong> barometric pressure<br />

and temperature seemed simple.<br />

The cost <strong>of</strong> 10 such buoys and a central processing facility was estimated<br />

to be about $750,000. Since this was more money than was available in the NSF<br />

contract for this purpose, outside financial support was needed. The obvious<br />

place to go was the NOM Data Buoy Office, and a visit was made in late 1972<br />

to explain the requirements. Subsequent events led to contracts by NDBO for<br />

concept definition and development and construction <strong>of</strong> three buoys, but not<br />

in time for field tests <strong>of</strong> a complete buoy prior to the main experiment. The<br />

<strong>AIDJEX</strong> <strong>of</strong>fice contracted for seven additional buoys and for the central station<br />

hardware. A considerable amount <strong>of</strong> testing was necessary on the ice at<br />

the beginning <strong>of</strong> the main experiment. Several refinements in the electronic<br />

and physical construction <strong>of</strong> the buoys were made on the ice under pressure to<br />

3


complete the deployment <strong>of</strong> the array. Eight <strong>of</strong> these buoys were in operation<br />

by late May 1975 and produced a spectacularly precise picture <strong>of</strong> ice movement.<br />

Unfortunately, the euphoria was short-lived,as half <strong>of</strong> the buoys quit<br />

in midsummer due to what was later found to be corrosion caused by<br />

moisture from a still unexplained source. During a brief period in the fall<br />

some <strong>of</strong> the failed buoys were recovered and one <strong>of</strong> these was put back into<br />

service. Four buoys continued to operate with little change through the balance<br />

<strong>of</strong> the experiment. As predicted, communication in the winter months<br />

was poor since it had not been possible to achieve reliable transmission <strong>of</strong><br />

data on the backup HF frequency. One <strong>of</strong> the two types <strong>of</strong> barometers employed<br />

in each buoy also was affected adversely by the moisture in the summer. As<br />

later calibrations revealed, the other type <strong>of</strong> barometer was unaffected and<br />

produced quite good pressures. These shortcomings might have been eliminated<br />

had there been time for field tests in the summer <strong>of</strong> 1974.<br />

Though uncertainties concerning the launch <strong>of</strong> the Nimbus 6 satellite<br />

precluded complete dependence on the spacecraft's Random Access Measurement<br />

System (RAMS) for buoy tracking and data relay, the relatively low cost <strong>of</strong><br />

RAMS transceivers made them attractive risks as a backup system for the HF-<br />

NavSat buoys. Originally, it was planned to include a RAMS transmitter on<br />

each <strong>of</strong> the HF-NavSat buoys, but a more attractive alternative was <strong>of</strong>fered<br />

when <strong>Polar</strong> Research Laboratory (PRL), with Office <strong>of</strong> Naval Research support,<br />

proposed construction <strong>of</strong> completely separate buoys using the RANS electronics<br />

and the barometers planned for use as backup sensors on the HF-NavSat buoys,<br />

and acoustic sensors provided by PRL. With logistic support incidental to<br />

the HF-NavSat buoy program, the RAMS buoys were placed adjacent to the larger<br />

buoys to provide an additional degree <strong>of</strong> security against loss or destruction.<br />

This arrangement proved to be crucial to the success <strong>of</strong> the experiment when<br />

the HF-NavSat buoy failures occurred.<br />

Even though the RAMS buoys had to be deployed prior to the launch <strong>of</strong><br />

Nimbus 6, eight <strong>of</strong> ten worked reliably and had marginally adequate position<br />

and pressure accuracy. During the experiment about two-thirds <strong>of</strong> the position<br />

data and about half <strong>of</strong> the pressure data from the main array were obtained<br />

from these buoys. The RAMS buoys were refurbished in 1976, and two continued<br />

in operation more than two years after original deployment.<br />

4


The scene during deployment <strong>of</strong> HF-NavSat buoy #3 at 77"12'N 161"43'W on<br />

9 May 1975. The superstructure rests on three battery tubes and an electronics<br />

tube, all penetrating through the ice into the water. The electronics<br />

tube is beneath the mast, behind the legs <strong>of</strong> the person in the<br />

center <strong>of</strong> the structure. A lead opened in 1976 where the people are<br />

standing in this picture. (Pat's picture)


HF-NavSat buoy 113 on 20 April 1976 at 76"44'N 16O0O2'W, the only visit<br />

to this site after deployment. The upper photo shows that structural and<br />

electrical connections to one battery tube failed as designed. The battery<br />

tube is intact in the lower photo to the left <strong>of</strong> the Twin Otter airplane,<br />

which landed between the buoy pieces on the refrozen lead. The buoy was<br />

found and left in good working condition. (Pat's pictures)


The initial <strong>AIDJEX</strong> scientific plan called for deployment <strong>of</strong> buoys to<br />

monitor ice motion in the shear zone or nearshore area. Cost considerations<br />

forced the elimination <strong>of</strong> these buoys in subsequent plans for the main experiment.<br />

Restoration <strong>of</strong> a portion <strong>of</strong> the nearshore buoy array was made possible<br />

under a new program to monitor nearshore conditions, which was funded by the<br />

OCSEAP <strong>of</strong> the Bureau <strong>of</strong> Land Management through NOM. The new program permitted<br />

two new developments in sea ice buoy technology: a buoy with current<br />

meters suspended in the mixed layer, in addition to pressure and temperature<br />

sensors (met-ocean buoys); and ADRAMS, a buoy which could be deployed by<br />

parachute froman aircraft and which operated on the surface <strong>of</strong> the ice rather<br />

than in a tube through the ice.<br />

Both development efforts were successful. The met-ocean buoys collected<br />

ocean current data including observations <strong>of</strong> inertial oscillations, propagating<br />

features, mean currents, and floe rotation which agree with and expand on<br />

observations taken at the manned stations. The air-droppable buoys (ADRAMS<br />

since they incorporate the RAMS) have had a remarkable record <strong>of</strong> reliability<br />

for such an early stage <strong>of</strong> development. About 50 (30 in <strong>AIDJEX</strong>) have been<br />

deployed by parachute without a failure, and the average operating life has<br />

met the design goal <strong>of</strong> 180 days. Equally important, the data from these buoys<br />

have added to the understanding <strong>of</strong> the ice movement near the shear zone and<br />

<strong>of</strong> the jet-like movement along the northwest Alaska coast. With the support<br />

<strong>of</strong> the NOAA Data Buoy Office and OCS, several ADRAMS have been fitted with<br />

barometers. Compensation for the wide temperature fluctuations <strong>of</strong> the ice<br />

surface environment has been very successful and several barometer-equipped<br />

ADRAMS have been deployed in the Chukchi Sea and Beaufort Sea. Most plans<br />

for future sea ice buoy work call for barometer-equipped air drop buoys.<br />

The total spent on data buoys during <strong>AIDJEX</strong> was about $1.7 million,<br />

$600,000 <strong>of</strong> it provided by NSF, $650,000 by NOAAjNDBO, $300,000 by OCS, and<br />

$140,000 by ONR.<br />

In this bulletin we review the data buoy and positioning system developments<br />

associated with the main <strong>AIDJEX</strong> experiment. Earlier work is covered in<br />

<strong>AIDJEX</strong> <strong>Bulletin</strong>s No. 7 and No. 22, available from NTTS. I appreciate the<br />

cooperation and help <strong>of</strong> all the people who made these technical initiatives<br />

possible.<br />

5


1. Introduction<br />

ARCTIC ODYSSEY--FIVE<br />

Pat Martin and C. R.<br />

YEARS OF DATA BUOYS<br />

Gillespie<br />

<strong>AIDJEX</strong><br />

4059 Roosevelt Way, N.E.<br />

Seattle, <strong>Washington</strong><br />

IN <strong>AIDJEX</strong><br />

The objective <strong>of</strong> the Arctic Ice Dynamics Joint Experiment (ALDJEX) is to reach a better understanding<br />

<strong>of</strong> the interaction <strong>of</strong> sea ice with the environment [l]. The main field experiment<br />

ended in Play 1976 after one year <strong>of</strong> data collection on the Arctic Ocean. Data buoys were used<br />

to define the motion <strong>of</strong> ice on the perimeter <strong>of</strong> the area <strong>of</strong> interest and to measure the surface<br />

barometric pressure over the same area (Fig. 1).<br />

2. Data Buoys<br />

2.1. Data Results<br />

A pilot experiment was conducted on the Arctic Ocean in the spring<strong>of</strong>1972, during which data<br />

buoys made detailed measurements <strong>of</strong> barometric pressure to develop and test atmospheric boundary<br />

layer theories in preparation for the main experiment. An array <strong>of</strong> buoys which operated<br />

for a year after this experiment provided drift and surface pressure and temperature data whic3<br />

were also used to plan the main experiment [2].<br />

During the 1975-1976 experiment, time series <strong>of</strong> position and pressure were obtained from a rix:<br />

<strong>of</strong> 6-8 sites with accuracies adequate for numerical modeling. Two months <strong>of</strong> these positiondars<br />

were accurate to 100'meters.<br />

An array <strong>of</strong> buoys with about 100-kilometer spacing along the coast <strong>of</strong> the Beaufort Sea providcl<br />

position data needed to interpret shore effects on the main array. These buoys should last<br />

through 1976 and, together with the buoys further from shore which should last until 1978, will<br />

provide data valuable to ice research and forecasting along the north coast <strong>of</strong> Alaska.<br />

2.2. Development Results<br />

Six types <strong>of</strong> sea ice buoys have been developed over a period <strong>of</strong> five years. Each has unique<br />

features which have met data collection requirements and contributed to the evolution <strong>of</strong> new<br />

designs. The principal characteristics <strong>of</strong> each are given in Table 1. Actual field experience<br />

with these buoys and the technologiecincorporated in them now totals about 30 buoy-years and<br />

is increasing at a rate which makes full evaluation <strong>of</strong> hardware performance and data very difficult<br />

(Fig. 2).<br />

-Six Arctic Data Buoys were deployed in 1972; four <strong>of</strong> these lasted about one year. The longesr<br />

life was 667 days and the longest drift track was 1800 kilometers. This was the first use <strong>of</strong><br />

a lowpower satellite link, the Interrogation, Recording and Location System (IRLS),forposit;icn<br />

and data retrieval. The spar configuration <strong>of</strong> this buoy was successful and has been used extensively<br />

since. Experience with this design led directly to later RAMS buoy use [3].<br />

Four <strong>of</strong> five Short Range Arctic Measuring Stations (SHRANS) operated 40 days in 1972 and provided<br />

a detailed record <strong>of</strong> barometric pressure which led to revised sampling on later buoys arJ<br />

to better field calibration procedures for barometers [4].<br />

Eight HF-WavSat buoys were deployed in 1975 and lasted for 60 days. Four <strong>of</strong> these lasted 11<br />

months. They raised the capabilities <strong>of</strong> sea ice buoys to a new level, with complete independ- ,<br />

ence from experimental satellite systems. They featured the only operational use <strong>of</strong> fullaccuracy<br />

Navy Navigation Satellite System fixes for data buoys. The high-frequency radio link<br />

was married to both the 3- and 12-day memories in the buoy to successfully overcomearctic radir,<br />

propagation conditions. A fully automated command, control and data reduction facility prove:<br />

its value in data collection over a full year. Data gaps in the winter emphasized the need fir<br />

an alternate communications frequency, which was disabled when the antenna could not be properLy<br />

tuned in the available time. Corrosive failures in the electronics point totheneed for better<br />

sealing from the humid summer sea ice environment. Two different pressure sensors were used<br />

and should make it easier to isolate errors [5].<br />

Eight <strong>of</strong> ten Synoptic RAMS (SYNRAMS) buoys deployed in 1975 before the Nimbus VI launch workei<br />

and were among the first buoys acquired when the Random Access Measurement System (RAMS) was<br />

activated. These buoys featured the first synoptic sampling and memory to work througha<br />

7


poizr-orbiting satellite, which is important where uniform samples are needed and satellite<br />

co~~erage is poor. This pointed the way toward fuller exploitation <strong>of</strong> the RAMS dnta capacity<br />

I6l-<br />

TWO <strong>of</strong> four Meteorological and Oceanographic (Met-Ocean) buoys operated for four months and one<br />

coiitinues to work after seven months. A 3Gsecond transmission interval coupled with platfor;?<br />

address switching extends the RAMS data capability to 64 bits per minute. Current speed and<br />

directlon at two levels in the mixed layer and magnetic buoy azimuth are measured.<br />

FiEteen Air-Droppable RAMS (ADRAMS) buoys have been deployed and all continue to operate after<br />

a maximum <strong>of</strong> six months. The small electronic and battery package designed to withstand the<br />

extreme Arctic surface environment, coupled with parachute deployment, has revolutionized ice<br />

buoy deployment logistics.<br />

3. Barometric Pressure Sensors.<br />

Recent progress in other areas has been so rapid that the major technical challenge to remote<br />

dnta collection is the need for better sensors. Our experience with barometers does not contradict<br />

that conclusion, Our use <strong>of</strong> pressure sensors is perhaps unique in that we are able to<br />

revisit buoys to perform calibration checks relatively easily 143. The preliminary resultsfor<br />

a small sample <strong>of</strong> the pressure sensors used in the past year show that the Hamilton Standard<br />

vibrating steel cylinder transducers drifted 0.1 millibar or.less during the year. There appears<br />

to have been drifts in the reported buoy pressures measured with the Paroscientific vibrating<br />

quartz beam transducers <strong>of</strong> from 1 to 4 millibars, but recalibration <strong>of</strong> the sensors in the laboratory<br />

shows drifts <strong>of</strong> from nil to 0.7 millibars. The cause <strong>of</strong> this discrepancy is not unders:ocd<br />

at this writing, but the possibility <strong>of</strong> an anomaly in the buoy electronics exists and<br />

Yhould be a reminder that good transducers do not guarantee good measurements.<br />

T!ie outputs <strong>of</strong> the barometers in use vary about 10% over full scale or about 1% over 100 millibars,<br />

Thus, 0.1 millibar is one part in lo5 <strong>of</strong> the transducer output. Clocks in buoys must<br />

56 better than one part in lo5 to avoid errors from this source. Also, synoptic sample timing<br />

requires an accuracy <strong>of</strong> one part in lo5 to be good to 5 minutes per year. Cursors in the data<br />

QT mode bits can permit calibration <strong>of</strong> the sample clock referenced to the satellite clock for<br />

synoptic systeas. Where real-time sampling is used, the buoy clock can still be calibrated if<br />

the buoy transmissions are controlled by the main buoy clock, including the unmodulated carrier<br />

por:ion <strong>of</strong> the transmission. This method <strong>of</strong> controlling transmissions may also have benefits<br />

in enhanced positioning schemes, and has been implemented on ADRAMS.<br />

4. MRS Position Fix Accuracy<br />

In order to provide a basis for the interpretation <strong>of</strong> the position fixes <strong>of</strong> the 25-30 drifting<br />

RAMS buoys now in use in the Arctic Ocean, a study <strong>of</strong> aMS position errors has been made.<br />

XL3)3EX receives data from NASA in the form <strong>of</strong> copies <strong>of</strong> the Nimbus archive magnetic tape with<br />

S.~ion resolution <strong>of</strong> the order centimeters and various supplementary information including<br />

satel;.ite position data and reference platform fixes.<br />

&,is 4long-Track Errors<br />

ExaEination <strong>of</strong> the fixes <strong>of</strong> RAMS platform 1337, located on St. Lawrence Island in Alaska, shows<br />

"1e radial error (68th percentile <strong>of</strong> the radial distribution) varies from 2 to 6 kilometers<br />

(Fig, 3). The RhVS system specification <strong>of</strong> 5 kilometers is met or exceeded for about 85% <strong>of</strong><br />

;he weeks analyzed over a six-month period. The error pattern <strong>of</strong> a reference platform operatel<br />

by KASA near Fairbanks is nearly identical to that <strong>of</strong> 1337, which suggests that the errors are<br />

from the same source; in this case, errors in the predicted satellite orbit. In fact, NASA has<br />

identified irregul-arities in the along-track motion <strong>of</strong> the satellite to be the major source <strong>of</strong><br />

Ehese correlated position errors, and the along-track error component <strong>of</strong> reference platform<br />

fixes is monitored on a daily basis to permit corrections to the orbit predictions. Since the<br />

present method <strong>of</strong> operation relies on operator analysis <strong>of</strong> error trends, detected errors only<br />

serve to improve the assumed orbit for future fix computations. The typical error growth and<br />

correction pro.cess occurs over periods <strong>of</strong> a few days, so that the weekly statistics reported<br />

here tend to average the worst events.<br />

It is possible to use information on position errors <strong>of</strong> reference platforms directly to correct<br />

oositions <strong>of</strong> moving platforms. This can be done either by adjusting the fix <strong>of</strong> a moving platfcrc!<br />

with the error in latitude and longitude for the same pass <strong>of</strong> a reference platform, or by<br />

reso!.ving the error into along-track and cross-track components prior to the adjustment. The<br />

latter method requires information on the satellite orbit which is not normally provided to<br />

xsers, but it does have the advantage that it is not affected by changes in the relationship<br />

8


etween the satellite sub-track and the latitude-longitude grid between the two platforms.<br />

Over a separation <strong>of</strong> 1000 kilometers both methods give essentially the same improvement and<br />

result in radial errors below 2 kilometers most <strong>of</strong> the time (Fig. 3).<br />

NASA could make an improvement in the positioning accuracy <strong>of</strong> RAMS, and save the user community<br />

a lot <strong>of</strong> extra work, if the along-track errors <strong>of</strong> fixed platforms were usedto correct the fixes<br />

<strong>of</strong> moving plat€orms. Similar arrangements should be made to improve Tiros N data, preferably<br />

by using reference platform information to enhance the orbit estimates, rather than by the<br />

direct adjustment <strong>of</strong> fixes. This was suggested for RAMS in a NASA-funded study in 1972 [8].<br />

__<br />

It is interesting that the radial position errors left after removal <strong>of</strong> along-track orbit errors<br />

show the NASA platform to be consistently better than 1337, which uses low costbouy electrazics<br />

and is exposed to large temperature fluctuations. There are significant periods <strong>of</strong> time whcr?<br />

the NASA platform radial error is at or below 1 kilometer, which is in good agreement with<br />

theoretical studies and which probably represents the best that can be done without enhancii;g<br />

the actual time and frequency measurements made in the spacecraft [9]. With high quality pktform<br />

hardware and implementation <strong>of</strong> s<strong>of</strong>tware modifications the positioning accuracy <strong>of</strong> RAMS<br />

could probably be brought to about 500 meters.<br />

4.2. RAMS Fix Editing<br />

The accuracy figures quoted above include all fixes reported by NASA withoutanyediting. Wkile<br />

correction for along-track errors is more effective than any edit scheme we have tested as protection<br />

against systematic errors, there are editing schemes which do seem to be selective L-<br />

eliminating bad data (Fig. 4). All editing schemes make a compromise between eliminating bat<br />

data and preserving as many fixes as possible. The quality index provided by NASA and the<br />

ber <strong>of</strong> messages used in the fix computation perform better than the other edits in this regzrd.<br />

After either <strong>of</strong> these edits about 80% <strong>of</strong> the fixes remain and the radial'error <strong>of</strong> the editel<br />

data is about 50% <strong>of</strong> the unedited errors when along-track errors are not eliminated, and abxt<br />

75% when along-track errors have been removed. Edits which might also prove effective when<br />

data from two passes are used in the fix, but which we have not tested, are those which detect<br />

unrealistic velocities or large differences between the two bias frequencies.<br />

About one-third <strong>of</strong> the fixes have larger errors than the numbers given here as the 68th percentile<br />

radial error statistic. If one has enough information about the limits <strong>of</strong> motiori <strong>of</strong><br />

the platform, then other editing schemes can be derived which eliminate the unreasonable date.<br />

In our case, since we get about 10 fixes per day for each platfom., and the ice usually moves<br />

less than 10 kilometers per day, the only editing used other than along-track error correcticn<br />

is a running median filter. A more sophisticated Kalman filter is then used to produce the<br />

final position and velocity estimates based on the RAMS fixes [lo].<br />

5. Assessment <strong>of</strong> <strong>Polar</strong> Satellite Data Collection and Tracking<br />

The primary advantage <strong>of</strong> data collection and tracking through a polar satellite system such As<br />

RAMS is platform simplicity. This feature makes possible high platform reliability with low<br />

cost. Those who are familiar with other more complicated satellite data collection and tracking<br />

systems can hardly mistake the message <strong>of</strong> widespread use <strong>of</strong> W S .<br />

The HF-NavSat buoy array cost about $800,000 more than the same spatial array <strong>of</strong> RAMS buoys 236<br />

was less reliable. If the HF-NavSat buoys had been as reliable as the EtAMS buoys, the cost cf<br />

a one-year experiment would have been about $250 per buoy-day versus $50 per buoy-day for t'kt<br />

RAMS buoys. The €IF-NavSat buoy positioning accuracy <strong>of</strong> 100 meters was valuable to <strong>AIDJEX</strong>; 3-i<br />

the compelling reason for these buoys was that we had to have data even if Nimbus VI didn't<br />

survive. Rams buoy simplicity is made possible by total rel&"e,on a satellite system whick<br />

is neither simple nor cheap. Users should familiarize themselves with the total system; not<br />

just the buoys and the data package received from NASA. The active participation<strong>of</strong> aninforLc3<br />

user community in full partnership with space organizations is essential to real progress.<br />

As buoy costs approach $10 per buoy-day, the cost <strong>of</strong> sensors (and their simplicity and reliability)<br />

becomes an important consideration. Taken one step further, the cost <strong>of</strong> data processiq<br />

becomes a major factor. These are signs <strong>of</strong> progress since no one is really after buoys or<br />

sensors, but data. Careful planning should include the cost <strong>of</strong> finished data. Data is the<br />

realcurrency and the exchange rate is improving.<br />

6. Satellite System Flexibility Concepts<br />

The random access approach <strong>of</strong>fers flexibility in data collection capacity and positioning accuracy<br />

at the user's option. A normal real-time sampling system transmits about 300 bits <strong>of</strong> data


t@ the spacecraft during each orbit. The spacecraft memory can be used as an extended buoy<br />

memory to permit the integration <strong>of</strong> multiple one-minute samples into ten-minute averages. i .-.e<br />

treatmcrit <strong>of</strong> each data reception at the spacecraft as independent allows the use <strong>of</strong> multiplc<br />

addresses and higher than nominal transmit duty cycles to obtain increased da ta capacity ..?Ti ::-.<br />

no increase in equipment costs. The availability <strong>of</strong> up to 13 sntelljte passes per dav in i)b-:z~<br />

regions means that up to 4000 bits per day can be transmitted for each address. There is<br />

trade-<strong>of</strong>f between the nrimber <strong>of</strong> data bits and level <strong>of</strong> redundancy required, and the nu::ber PI<br />

satellite passes available and number <strong>of</strong> addresses used for each platform. This trade-cif I :-<br />

a function <strong>of</strong> latitude, and with the number <strong>of</strong> passes available in polar regions we hi17.e h<br />

extremely conservative with respect to data redundancy in our system designs. This should ch2:;e<br />

as we become more confident <strong>of</strong> system performance.<br />

Flexibility in position accuracy requires that measurements made in the satellite have sufficient<br />

accuracy potential for the most demanding users. To achieve refraction-limited posit<br />

accuracy <strong>of</strong> 200 meters, a Doppler system like RAYS needs measurement accuracy <strong>of</strong> 0.1 Hz and<br />

0.61 seconds, an order <strong>of</strong> magnitude better than RA". The use <strong>of</strong> multiple addresses <strong>of</strong>fers<br />

the possibility <strong>of</strong> increasing the amount <strong>of</strong> data used in the position computation, as a way :f<br />

improving fix accuracy. The flexibility <strong>of</strong>fered to the user by random access data collecticx<br />

and positioning systems should be retained and enhanced in the future.<br />

7. Real-Time Data Readout Option<br />

The Ni.mbus spacecraft transmits all data in real time in addition to storing these data on rzznetic<br />

tape. When the spacecraft is critiiin view the data can be receivedwit'nportable equip.-::t<br />

and, with the proper formats, can be decoded and interpreted essentially the instant they arc<br />

transmitted from the platform. The most obvious use <strong>of</strong> this capability for data buoy work is<br />

to provide for inmediate verification <strong>of</strong> platform performance during checkout <strong>of</strong> electrcnics.<br />

This is the only practical method <strong>of</strong> authoritative confirmation <strong>of</strong> buoy pcrfor;nance in remozt<br />

areas.<br />

The same capability could be used to obtain data for forecasting without waiting for process5r;g<br />

by NASA.<br />

The primary purpose <strong>of</strong> this real-time data transmission is to provide a backup data collect<br />

mode in case <strong>of</strong> tape recorder failure in the spacecraft. The spacecraft must be in view <strong>of</strong><br />

both the platforms and the data retrieval site for sufficient time to recover the desired dzr3,<br />

Thus, the data retrieval site should be in the vicinity <strong>of</strong> the platforms, which is <strong>of</strong> specic:<br />

significance to experiments in remote areas where NASA does not maintain facilities. The us?<br />

<strong>of</strong> buoys in the Southern Ocean might be one experiment which would benefit from a direct re;;-<br />

QUt capability in the Antarctic in the event <strong>of</strong> a tape recorder failure.<br />

An important addition to the real-time data transmission <strong>of</strong> the spacecraft would be the predicted<br />

orbit parameters and measured satellite clock and orbit errors needed for position fi-:<br />

computation. This would permit fix computation by users and wou1.d make possible dispersing<br />

the data processing function around the globe to areas where the data are being collected ar-i<br />

used.<br />

8. Conclusion<br />

The data buoy program associated with <strong>AIDJEX</strong> has been large and diverse enoughtogainanapprsciation<br />

for the many practical difficulties involved in developing and using several kinds CT<br />

data buoys. Time seems to have been the ingredient in shortest supply, and should t?,ereforn<br />

be canserved in future work. The money spent (about $2 million dollars including logistics<br />

costs), while not extravagant, must certainly be considered adequate for the objectives, whi:i<br />

were fairly well sati.sfied. Personal initiative was important, and more <strong>of</strong> this ingredient<br />

would have improved the results.<br />

We have been fortunate to try many new things with very few failures, but this should not bs<br />

interpreted as an endorsement <strong>of</strong> the policy <strong>of</strong> trying something new when a proven method car.<br />

be made to work. The only good pressure sensor is a used one. This is a good philosophy frz<br />

buoys as well, althoush one probably has to settle for used designs and long checkout tests<br />

since hardware recovery is difficult.<br />

The RAMS has 2 to 6 kilometer position accuracy and is probably capable <strong>of</strong> 500 meters. By<br />

taking full advantage <strong>of</strong> the flexibility available through random access, a data capacity <strong>of</strong><br />

one to several thousand bits per platform per day can be acheived. There are real benefits 10<br />

be achieved through real-time readout <strong>of</strong> data from the spacecraft. These proven captbilities<br />

<strong>of</strong> FUtIS, plus the potential ready to be realized, make it highly desirable that a new<br />

I<br />

10


deployment <strong>of</strong> the RAMS be scheduled so that the system can be utilized through the rest <strong>of</strong> :his<br />

decade. A new orbit complementary in time to the existing one would be desirable.<br />

Data collection and tracking from orbiting spacecraft is in its infancy. By comparison the<br />

Navy Navigation Satellite System (Transit) has been an operational system for 14 years, and<br />

currently consists <strong>of</strong> six Satellites which provide users with a completely self-contained pcsitioning<br />

capability. The insight gained in the use <strong>of</strong> this system has lead to a new satellite<br />

system which will provide continuous position information to a few meters, from a constellarion<br />

<strong>of</strong> up to 24 satellites. The message is not that we need 24 data collection satellites. Ths<br />

current concept <strong>of</strong> data collection and tracking is elegant in its simplicity. The challeng-. is<br />

to raise our level <strong>of</strong> sophisitication in using the data collection and tracking concept tohigher<br />

levels where even greater benefits will be realized.<br />

9. Acknowledgments<br />

This work was supported by the National <strong>Science</strong> Foundation, the Office <strong>of</strong> Naval Research, the<br />

National Oceanic and Atmospheric Administration, the Bureau <strong>of</strong> Land Management, the Canadiaii<br />

<strong>Polar</strong> Continental Shelf Project, and the National Aeronautics and Space Administration. We<br />

wish to thank the staffs <strong>of</strong> the Nimbus Operations <strong>Center</strong>, Goddard Space Flight <strong>Center</strong>; <strong>Polar</strong><br />

Research Laboratory, Santa Barbara; the Applied Physics Laboratory, <strong>University</strong> <strong>of</strong> Washingtoz,<br />

and our colleagues at <strong>AIDJEX</strong>.<br />

10. References<br />

[l] N. UNTERSTEINER, <strong>AIDJEX</strong> <strong>Bulletin</strong> 26<br />

[2] P. MARTIN, in: Means <strong>of</strong> Acquisition and Communication <strong>of</strong> Ocean Data, WMO No. 350, 1973.<br />

[3] D. P. HAUGEN and K. M. DOZIER, Applied Physics Laboratory, <strong>University</strong> <strong>of</strong> <strong>Washington</strong>,<br />

APL-UW 7422 (1975).<br />

[4] W. P. BROWN, <strong>AIDJEX</strong> <strong>Bulletin</strong> 22 (1973).<br />

(1974).<br />

[5] W. P. BROWN and E. G. KERUT, Ocean 75, IEEE Publication 75 CHO 995-1OEC (1975).<br />

[6] S. P. BURKE and B. M. BUCK, Ocean 75, IEEE Publication 75 CHO 995-1 OEC (1975).<br />

[7] P. MARTIN, Ocean 74, IEEE Publication 74 CHO 873-0 OCC (1974).<br />

[8] Nimbus F TWERLE Doppler Data Processing, General Electric, Space Division No. 72SD4257<br />

(1972).<br />

[9] T. GREEN, Geoscience Electronics, GE-13, No. 1 (1975).<br />

[lo] A. S. THORNDIKE, <strong>AIDJEX</strong> <strong>Bulletin</strong> 24 (1974).<br />

11


BUOY NAHE<br />

COMMUN. COMMANDS FIXES<br />

TABLE 1. DATA BUOY CHARACTERISTICS<br />

' NOMINAL<br />

BITSIDAY X BATTERY TYPE NO. BUILT/<br />

SAMPLING REDUNDANCY SENSORS f CAPACITY STRUCTURE TOTAL COST<br />

5<br />

A<br />

k<br />

rt<br />

ri<br />

P,<br />

Arctic<br />

Data Buoy<br />

IUS<br />

Nimbus<br />

Beacon 2-5 km Real time<br />

turn on IRLS 6 orblday<br />

Rangerange<br />

343 x 3 Pressure<br />

Temperature<br />

Voltage<br />

Mercury<br />

Polyethylene<br />

3 KWH Spar buoy<br />

7<br />

$150.000<br />

SBRAMS<br />

VHF<br />

Aircraft<br />

Transmit None<br />

data<br />

Hourly 240 x 2 Pressure<br />

24 orblday<br />

Lead Acid PVC tube<br />

0.7 KWH external<br />

antenna +<br />

battery<br />

7<br />

$100 000<br />

P<br />

N<br />

HF-N avSat<br />

HF- 4HZ<br />

Beacon 100 meters 3 Hourly 1632 x 8<br />

transmit NavSat 8 orblday + Colaaand<br />

+ sample Doppler + Command<br />

clock<br />

shift<br />

Pressure<br />

Temperature<br />

Volt age<br />

Position<br />

Zinc-Carbon-Air<br />

Polyethylene<br />

+ Lead Acid + Aluminum<br />

45 KWH 4-leg spar<br />

buoy with<br />

aluminum upper<br />

s truc tute<br />

10<br />

$1 0oo.Ooo<br />

SYNRAMS<br />

RAMS<br />

Nimbus<br />

None<br />

1.5-5 ken<br />

RAMS<br />

Doppler<br />

3 Hourly<br />

8 orblday<br />

256 X 15<br />

Pressure<br />

Temperature<br />

Ambient noise<br />

Zinc-Carbon-Air Aluminum spar<br />

1 KMH buoy<br />

11<br />

$140 000<br />

Me t-Ocean<br />

RAMS<br />

Nimbus<br />

None<br />

1.5-5 ~UI<br />

RAMS<br />

Doppler<br />

3 Hourly<br />

8 orblday<br />

512 x 15<br />

Pressure<br />

Temperature<br />

Voltage<br />

Azimuth<br />

Current speed<br />

+ Direction<br />

Zinc-Carbon-Air Polyethylene<br />

1 KWH spar buoy<br />

4<br />

$104 000<br />

ADW<br />

m<br />

RAMS<br />

Nimbus<br />

None<br />

1.5-5 km<br />

RAHS<br />

Doppler<br />

Real t b e<br />

10-12 orb<br />

f day<br />

352 x 11 2 buoys with<br />

Pressure<br />

Temperature<br />

Inorganic<br />

Lithium<br />

1.5 Kwtl<br />

Lexan<br />

sphere<br />

17<br />

$192 000


(Martin & Gillespie)<br />

Fig. 1. Initial and final array <strong>of</strong> buoys for <strong>AIDJEX</strong> main experiment,<br />

June 1975-May 1976: 0-0, initial array; 0-0, final . _ _ "<br />

array;, 0 , additional buoys.<br />

20<br />

.<br />

15<br />

Estimate<br />

5<br />

* I<br />

I<br />

I<br />

i<br />

'71 '72 ' 73 '74 ' 75 '76<br />

Year<br />

Fig. 2. Accumulation <strong>of</strong> data from buoys during <strong>AIDJEX</strong>, 1971-1976.<br />

13


L(<br />

0 I 2 3 4 3<br />

Radial error, kilometers<br />

-/.<br />

?ig. 3. Effect <strong>of</strong> various editing techniques on ?AIS accuracy with<br />

and without along-track errors for NASA reference platform, 15-<br />

20 Feb. 1976. Lower case--along-track errors are present: upper<br />

case--along-track errors are removed. Type <strong>of</strong> edit--8 = none;<br />

A = elevation angle; B = messages used > 12- C = messapes used<br />

> 8; T = two pass fixes only; F = A+T; Q = quality inclex .? 47.<br />

H = A+T+q.<br />

r l<br />

i-? ~<br />

Improvement in TMS ~osition errors by removtnp ?Xong-tracL<br />

satellite errors. IB platform 1337, mcorrected: t platforr; 1237.<br />

corrected; NASA platfom,, corrected,<br />

14


Arctic Environmental Buoy System<br />

Walter P. Brown<br />

<strong>Polar</strong> Research Lab., Inc.<br />

Edmund G. Kerut,<br />

NOAA Data Buoy Office<br />

ABSTRACT<br />

The AEB is a remote unattended data acquisition<br />

and telemetry system designed for deployment<br />

on ice covered seas. The total system as<br />

presently configured consists <strong>of</strong> up to 12 AEBs<br />

and a Central Control Station (CCS). The Central<br />

Control Station under computer control<br />

collects the data from the AEBs, processes the<br />

data and formats the data on a digital tape for<br />

future analysis. The CCS is also capable <strong>of</strong><br />

controlling the majority <strong>of</strong> the AEB functions<br />

via a command link. The AEB is configured to<br />

sample sensor data and acquire position data<br />

at three hour intervals automatically. The<br />

present sensor configuration allows 6 primary<br />

sensors with 10 bit resolution and 16 auxiliary<br />

sensors with 5 bit resolution. The auxiliary<br />

sensors are sampled only once per day. The<br />

sensor data and position data are stored in a<br />

digital memory which is transmitted via an H.F.<br />

link once per day to the Central Control Station.<br />

A unique dual memory concept is utilized<br />

to prevent data loss due to propagation vagaries<br />

and polar cap absorption events. The<br />

position measurements are accomplished by an<br />

on-board NAVSAT receiver.<br />

INTRODUCTION<br />

The NOM Data Buoy Office (NDBO) engineering<br />

development activities include the development<br />

<strong>of</strong> arctic data buoys in support <strong>of</strong> national and<br />

international scientific experiments. As part<br />

<strong>of</strong> these activities a program has recently been<br />

successfully concluded to develop and test three<br />

prototype arctic environmental buoys (AEB) to<br />

provide the remote data requirements <strong>of</strong> a<br />

scientific experiment designed by the kctic<br />

- Ice gynamics Joint Eperiment (<strong>AIDJEX</strong>) Project<br />

Office. The experiment is designed to investigate<br />

the large scale response <strong>of</strong> sea ice to<br />

changing environmental parameters. The <strong>AIDJEX</strong><br />

program as presently envisioned is the first <strong>of</strong><br />

a series <strong>of</strong> studies that will subsequently be<br />

incorporated under a large E a r seriment<br />

(POLEX). The objective <strong>of</strong> the <strong>AIDJEX</strong> experiment<br />

is to reach, through coordinated field experiments<br />

and theoretical analysis, a fundamental<br />

understanding <strong>of</strong> the dynamic and thermodynamic<br />

interaction between arctic sea ice and its environment<br />

and to answer basic questions <strong>of</strong> the<br />

mechanisms which cause large scale ice deformation<br />

and the effect <strong>of</strong> ice deformation and<br />

morphology on the heat balance.<br />

The experimental design requires the establishment<br />

<strong>of</strong> an array <strong>of</strong> drifting ice buoys in the<br />

Arctic Ocean to measure atmospheric pressure and<br />

temperature at the sea ice surface. An essential<br />

requirement <strong>of</strong> the buoy design was to develop a<br />

position determination capability, an order <strong>of</strong><br />

magnitude beyond the capability <strong>of</strong> polar orbiting<br />

meteorology satellites with position fixing capability,<br />

which would be operational during the<br />

experiment .<br />

The program phases included the following<br />

elements: the study <strong>of</strong> the experimental design<br />

for the <strong>AIDJEX</strong> experiment which involved the<br />

array <strong>of</strong> Arctic Data Buoys (AEB) and the translation<br />

<strong>of</strong> the system measurement requirements<br />

into a system specification; the design and<br />

development <strong>of</strong> prototype system hardware to meet<br />

the measurement requirements and test objectives<br />

<strong>of</strong> the AXDJEX experiment; fabrication <strong>of</strong> three<br />

prototype AEB's and associated test set for test<br />

and evaluation in the field prior to the main<br />

experiment; and performance <strong>of</strong> laboratory and<br />

field testing on the prototype system to verify<br />

experimentally its ability to meet the measurement<br />

requirements and test objectives <strong>of</strong> the<br />

main experiment. The design, development and<br />

fabrication program was performed by the <strong>Polar</strong><br />

Research Laboratory in Santa Barbara, California<br />

under contract to NDBO.<br />

DESIGN CONSIDERATIONS<br />

The conceptual design <strong>of</strong> the Arctic Environmental<br />

Buoy (AEB) System was formulated to meet<br />

both the near term requirements <strong>of</strong> the Arctic Ice<br />

Dynamics Joint Experiment (<strong>AIDJEX</strong>) and the general<br />

need for gathering data on the Arctic ice<br />

pack. The basic requirements were to sample a<br />

number <strong>of</strong> sensors on a synoptic basis (i.e.,<br />

every 3 hours starting at 0000 Zulu), provide an<br />

accurate position for the system 8 times a day<br />

and to transmit the data at least once per day.<br />

In addition the AEB system was to have an unattcnded<br />

life <strong>of</strong> 8-14 months. The range <strong>of</strong> the remote<br />

buoy stations from the Central Station is expected<br />

to be 250 to 500 Km during the life <strong>of</strong> the<br />

experiment and therefore an H.F. link was selected.<br />

Frequency selection and modulation methods<br />

were chosen on the basis <strong>of</strong> computer analysis<br />

and studies (1)(2)(3) performed at the Institute<br />

<strong>of</strong> Telecommunication <strong>Science</strong>s in Boulder, Colorado.<br />

The structural design <strong>of</strong> the buoy hull and<br />

antenna was heavily influenced by the unique<br />

environmental conditions <strong>of</strong> the Arctic ice pack.<br />

50 - IEEE OCEAN '75<br />

15


The buoy will be subjected to temperaebres as low<br />

as -5OOC in the winter with ictemit- ng<br />

and temperatures <strong>of</strong> + 5OC in tbe rumer w<br />

moist air and the possibility af flee floating,<br />

In addition the structure must be able LG dthstand<br />

the loving attentions <strong>of</strong> a polarr iear,<br />

The Navy Navigation Satellite (NAVSA<br />

selected for positioning becat'se o_C<br />

ly good accuracy (sjometers radius CEP, sirqie<br />

channel, co-location mode) and tPac eacc posf.tio~~<br />

fix defines a unique position ma !s ?cr depe?-<br />

dent on past history (lane counting) a8 is<br />

required in OMEGA or Global positim systems.<br />

The emphasis in the design was piaced os conserving<br />

power and weight and in ~rovidhg redmdancy<br />

in high risk areas. The volume ai.d weight<br />

<strong>of</strong> the system had to be compatible ?dth tFe<br />

capabilities <strong>of</strong> a Twin Otter aircrait.<br />

AEB SYSTEM DESCRIPTIOE<br />

The AEB system consists <strong>of</strong> 8 remote statiovs<br />

installed in a circular pattern with a radics ot<br />

300 to 400 kilometers from the central ALDJEX ice<br />

camp. These AEB's are received and interrogated<br />

by a Central Control Station (CC3) locaze? st the<br />

central <strong>AIDJEX</strong> ice camp. The AEB is a self contained<br />

unmanned telemetry statior. installed in<br />

the ice pack with a battery ?over suppLy capable<br />

<strong>of</strong> 14 months <strong>of</strong> continuous operatior., As presently<br />

configured the AEB's sample 60 bits (if<br />

envirorxnental data and 144 bits <strong>of</strong> pcsitior<br />

information every three hours. Addit'iora!.<br />

environmental and position samples caq 30 c0-f<br />

manded from the Central Control Station. R e<br />

AEB environmental sensor suite for AXDJEY consists<br />

<strong>of</strong> two atmospheric pressure senccrs and<br />

two temperature sensors, eacb. u~ing a 10 bi:<br />

word, thus leaving 2 spare 10 bit sensor inputs<br />

available.<br />

The position samples are obtained frob6 Nav'-<br />

gation Satellite receiver uhicl- provides GIE 24<br />

bit word <strong>of</strong> identification data and fd7;e 24 blt<br />

words <strong>of</strong> doppler data. Each <strong>of</strong> the sens3r and<br />

position word groups has 48 bits <strong>of</strong> sync an?<br />

reference information attached which ~c,~des a<br />

5 bit sample <strong>of</strong> engineering data. The evvlrrxmental<br />

data, engineering data and positLop data<br />

are stored in two digital IC m~mories. 9-e maory<br />

provides short term storage .sf 2 de 7c, "ata<br />

and a long term memory provides 12 days. <strong>of</strong> stcrage.<br />

The long term memory is used to c~vsr<br />

communication link interruptions <strong>of</strong> t~r TO 12 Says<br />

and to fill in holes fn daily transmissions. Two<br />

cycles <strong>of</strong> the short term memory axe era?snit:ed<br />

once per day and two cycles ot the lcng term<br />

memory once per 10 days on an automatic basis.<br />

Additional transmissions from eitber n%orv can<br />

be commanded by the Central Contrct S"bat"o- The<br />

AEB also contains a strobe light acZ 'XF beacon<br />

transmitter as location aids. Tbes? ~n;e_s Ere<br />

turned on automatically once per 13 days FoS<br />

three hours and can also be turned ca by coaaand<br />

from the CCS.<br />

The Central Control Station prev%des for data<br />

reception and remote control <strong>of</strong> the AZB's. Tt<br />

contains a Navigation Satellite -eceivnr ;rnic\<br />

is used to obtain local fixes and pr-v'des :\e<br />

additional message data necessary to calc~late<br />

M1B -osi:iccs. Received data is processed by e<br />

Rous 2,i'lC computer whjcb mssages the raw AEB<br />

sensor data into a finishel form and stores this<br />

iafcrmation on a digital magnetic tape. Th2 conpriter<br />

also provides control <strong>of</strong> the CCS operation<br />

arid sends commands tc the AES's as required.<br />

Srck-up Yecsrding <strong>of</strong> data is provided, in cese <strong>of</strong><br />

compxter faiiure. Ar. ,.ninterrL?tzhie power su?p€y<br />

furnishes primary power to criticzl components so<br />

chat data 5s not I=st d:;e :o t+,e frequent c,itPBes<br />

:~f ice CZR~ power.<br />

An Mi3 test set conc-aLns the ne-.essary eouipment<br />

to test and isolate problems in the A9B to<br />

aub-system level. Th.e.test set is also capable<br />

<strong>of</strong> sendir.8 comands te and receiving data from<br />

the AEB, thus simulating the CCS station.<br />

AEB ELPCTRONIC DESIGN<br />

W block liagrao <strong>of</strong> the AEB is shown in Figure<br />

1, A11 sensors except the engineering sensors<br />

are sampled qncc every 3 hours beginning at 0000<br />

hours Zulu. Starting at the top, the NAVSAT receiver<br />

is turned on €or one hour by the control<br />

electronics. The receiver begins sweeping and<br />

evectually locks onto a satellite. It then<br />

"Lacks the satellite until message synchronization<br />

is established. At this point it reqxests<br />

the control electronics to send it the Zulu time<br />

<strong>of</strong> day which it stores in an internal register.<br />

Alsa it extracts the relative Zulx time and the<br />

satellite I.D. Erom the -eceived message and<br />

stor~s this infomation. It then continues<br />

':racking che saifllite imtil five 2 minute<br />

d.oppler intervals are received. At this point.<br />

the NAVSAT receiver signals the memory and conrroP<br />

sub-systeas that data is rcady. At :he<br />

prcpes point ir. the memory cycle the data is<br />

transferred from the NAVSAT register into the<br />

memory. The ahcve description describes a normal<br />

NAVSAT cycle. Additional logic within the NAVSAT<br />

receivers allows for 2 other conditions. In one<br />

case the receiver could lose RF lock on the<br />

satellite before it had acquired 4 doppler neasuremenfs.<br />

Uilde~ this condition if the one hour<br />

gate .~K-!E the cortrol electronics is sti.11 high<br />

the NAVSAT .receiver will begin swceying again<br />

an? attempt to acquire another satellite. Ir.<br />

s~.oti:er case if four dopplers ar<br />

2.F lcck :s iost t:-e o7zrztlon wi<br />

as in :he :icxzaP case and cn1.y 4<br />

be stored whicn is sufficient for a position fix.<br />

The time thc NAVSAT receiver is ON will vary<br />

from 18.5 minutes to 1 hour. In general. hecausz<br />

nf the high number 06 satellite passes in<br />

thz Arctic the ON tiimr. should be less than 70<br />

minutes on the average.<br />

-.<br />

;ne yressure se2sors are turned<br />

a?.riutes at each s;moptic time. The<br />

zf rb? eteven IS used to allow -he sensors to<br />

stabilize, durlng the next PO minutes the outputs<br />

OS %he pressu~e transducers are averaged. The<br />

technique used is eo continuously count the<br />

freqxency output cf the pressure trsnsducers in<br />

a 10 bit counter and store the count at the end<br />

<strong>of</strong> the 19 minute period. Because the barometric<br />

pressure range <strong>of</strong> intcrest (SSO-lOSG nillibars)<br />

Ls only Z~GV,: a 1/10 ?f the tranducer pressure<br />

range, a suitable 6igital divider must be used


etween the output <strong>of</strong> the transducer and the 10<br />

bit counter to avoid an overload problem. With<br />

the 10 bit counter the barometric pressure resolution<br />

is a nominal .1 millibar.<br />

The two temperature sensors, the spare sensors,<br />

and the engineering data sensors are all conditioned<br />

to provide an analog output <strong>of</strong> 0 to -5<br />

volts. The temperature sensors cover a range <strong>of</strong><br />

-50°C to +lO°C with a resolution <strong>of</strong> .06OC. These<br />

outputs are multiplexed into a 10 bit AID converter<br />

and except for the engineering data sensors<br />

are dumped into the memory at each three<br />

hour sample period. The engineering data sensors<br />

are sampled only once per day with two <strong>of</strong> the<br />

sensors being entered into the memory each three<br />

hour sample period. Only the 5 most significant<br />

bits <strong>of</strong> the AID are used for the engineering data.<br />

The engineering data presently being measured are:<br />

RF power out and reflected power; primary and<br />

secondary battery voltages; several temperatures<br />

in the electronics housing and a leak detector.<br />

The short and long term memory are operated<br />

identically except that the long term memory is<br />

exactly 4 times longer than the short term which<br />

contains 7200 bits <strong>of</strong> storage. Both memories<br />

use dynamic MOS shift register chips connected<br />

in a serial recirculating memory configuration.<br />

Two types <strong>of</strong> words are entered into the memory;<br />

a sensor word which is 108 bits and a NAVSAT<br />

word which is 192 bits. The words are entered<br />

sequentially and once the memory is filled the<br />

new data replaces the oldest. The word formats<br />

are shown in Table 1. Both the sensor and<br />

NAVSAT words start with a sync pattern. This<br />

approach uses more memory space for non-data<br />

and requires more transmit time than a less<br />

frequent sync pattern approach such as once per<br />

day. However, it has the advantage <strong>of</strong> allowing<br />

sensor and NAVSAT words to be entered randomly<br />

with no word limit per day. It also has the<br />

advantage on the receiving end <strong>of</strong> improving the<br />

amount <strong>of</strong> data received. Fading is quite prominent<br />

on H.F. links and if a fade occurs during<br />

a sync pattern with this scheme only one'data<br />

word is lost where with a less frequeny sync<br />

pattern approach a larger block would be lost.<br />

The short term memory is normally connected to<br />

the Bi-phase L modulator and is transmitted on<br />

a daily basis. Once per 10 days the long term<br />

memory is switched into the modulator and transmitted.<br />

Two VCTCXO's are supplied to provide<br />

redundancy or the flexibility <strong>of</strong> dual frequency<br />

operation if needed. The present plans call for<br />

operation on a single frequency, therefore both<br />

VCTCXO's are the same. In the event <strong>of</strong> a failure<br />

<strong>of</strong> the VCTCXO being used, a command can be sent<br />

from the CCS station to switch to the alternate<br />

one. The VCTCXO's are passively combined and<br />

either can drive the 100 watt power amp. The<br />

power amp is configured such that either half<br />

can fail and still allow degraded communication<br />

with 25 watts output. The power amp is broadband<br />

and will provide its full output over a frequency<br />

range <strong>of</strong> 2 to 12 MHZ. A coax switch is used to<br />

switch the antenna from the command receiver to<br />

the transmitter during the transmit cycle. A<br />

matching network is used at the base <strong>of</strong> the<br />

sleeve dipole antenna to allow adjustment for<br />

various ice thickness. The command receiver<br />

operates on two frequencies to allow 24 hour<br />

command coverage. The present assignments are<br />

4.165300 and 2.146 MHZ. The control circuits<br />

switch the receiver between these two frequencies<br />

on a one minute cycle. This procedure eliminates<br />

the need for two separate receivers and is acceptable<br />

since none <strong>of</strong> the commands require immediate<br />

action. To assure reception <strong>of</strong> the command<br />

on the right frequency, the CCS station merely<br />

sends the command sequence twice with a one minute<br />

spacing, The command is decoded and sets up<br />

the required action in the control electronics.<br />

Two location aid devices are provided. A 300<br />

milliwatt VHF beacon on 108.1 MHZ allows an aircraft<br />

to "home" on the buoy from a range <strong>of</strong> 30<br />

to 50 miles. A strobe light allows visual sighting<br />

<strong>of</strong> the buoy in twilight or dark conditions<br />

with ranges up to 10 miles.<br />

The AEB power supply consists <strong>of</strong> three banks<br />

<strong>of</strong> primary carbon-air cells with each bank consisting<br />

<strong>of</strong> fifteen 1000 amp hour cells. Because<br />

<strong>of</strong> the nature <strong>of</strong> these primary batteries, they<br />

are not capable <strong>of</strong> providing the peak current<br />

demands <strong>of</strong> the system, therefore they are used<br />

to charge a secondary battery bank. The secondary<br />

battery consists <strong>of</strong> 36 Gates sealed lead<br />

acid cells arranged in three 12 volt banks which<br />

provide approximately 24 amp hours when fully<br />

charged at O°C. Individual chargers are used<br />

between the secondary battery banks and the<br />

primary banks and the secondary banks are diode<br />

isolated from the power buss. Thus, a failure<br />

in any <strong>of</strong> the banks will not stiut the system<br />

down but will reduce the life <strong>of</strong> the system.<br />

The master timing for the AEB is derived from<br />

a very stable oven controlled 5 MHZ oscillator.<br />

Typical stabilities for the oscillator are 1 X<br />

10-9 per 30 days. In 14 months there are 10,224<br />

hours, thus the error in time at the end <strong>of</strong> the<br />

experiment using this oscillator should be less<br />

than 1.0 seconds assuming that the above stability<br />

is a linear change over the 14 month life<br />

and that the frequency was properly set initially.<br />

AEB STRUCTURAL DESIGN<br />

The structure <strong>of</strong> the AEB as installed in the<br />

Arctic ice pack is shown in Figure 2. In developing<br />

the design full advantage was taken <strong>of</strong> some<br />

<strong>of</strong> the unique characteristics <strong>of</strong> the ice cover<br />

sea. The ocean water below the ice remains thermally<br />

stable with only 2OC variation and a mid<br />

point near O°C over the entire year (4). The<br />

ice itself acts as an insulator against the surface<br />

temperature extremes. Equipment installed<br />

under the surface <strong>of</strong> the ice will never see<br />

tehperatures lower than -2OOC even though the<br />

surface temperature may reach -5OOC (5), (6).<br />

These facts are utilized in the design by locating<br />

all the electronics and the batteries below<br />

the surface <strong>of</strong> the ice in 8" diameter tubes.<br />

The electronics modules are located in the central<br />

tube. The temperature sensitive components<br />

such as the master oscillator and barometers are<br />

located in the bottom <strong>of</strong> the tube which is surrounded<br />

by the sea water. The other electronics<br />

modules are placed above these in order <strong>of</strong> decreasing<br />

temperature sensitivity. The electro-<br />

52 - IEEE OCEAN '75 17


nics occupy about 12 feet <strong>of</strong> the 17.5 foot long<br />

tube which extends 3 feet above the ice, Thus<br />

the electronics are all below the surface. The<br />

tubes extend 3 feet above the ice to prevent<br />

possible flooding from melt ponds which xcur on<br />

the ice surface in the summer. The primary batteries<br />

are located in the outer 3 tubes and<br />

occupy approximately 14 feet <strong>of</strong> vertical space<br />

thus the batteries are also below the surface <strong>of</strong><br />

the ice. These tubes are the support for the<br />

upper structure and provide enough buoyancy to<br />

maintain the system at approximately the same<br />

level even in a full melt condition.<br />

The upper structure consists <strong>of</strong> 6 twelve foot<br />

radials with their supporting arms; a 6 inch diameter<br />

12 foot long sleeve which provides the<br />

mounting structure for the temperature sensors,<br />

the VHF Beacon antenna, the NAVSAT antenna, the<br />

strobe light and acts as the feed line for the<br />

H.F. whip antenna. The H.F. whip is 24 feet long<br />

and is center loaded for the 4.1653 MHZ transmit<br />

frequency. The abbreviated radial system was<br />

used in an attempt to minimize the effect on<br />

antenna impedance <strong>of</strong> possible shifts in the ground<br />

plane due to summer melt condition. The chains<br />

at the ends <strong>of</strong> the radials are used to dampen<br />

wind-induced vibration. The antenna structure<br />

was grounded to sea water through the battery<br />

tubes. The temperature sensors are located<br />

nominally at 1 meter and 3 meters above the ice<br />

surface in radiation shields attached to the<br />

sleeve. The strobe light, NAVSAT antenna and<br />

VHF Beacon antenna are spaced120' at the top <strong>of</strong><br />

the sleeve to minimize structural interference<br />

effects and to improve line <strong>of</strong> sight range.<br />

DEPLOYMENT<br />

The AEB's were deployed using a Twin Otter<br />

fixed wing aircraft equipped with skis. A multiyear<br />

ice flow in the desired location was selected<br />

and reconoitered for a suitable landing site.<br />

After landing,small pilot holes were drilled to<br />

determine ice thickness. If a suitable thickness<br />

<strong>of</strong> 8 to 12 feet were found, the installation was<br />

begun. Holes were drilled thru the ice for the<br />

four 8" diameter tubes using a 9" ice drill.<br />

The tubes were installed and the primary batteries<br />

loaded into the outer tubes using a tripod.<br />

These primary batteries could not be pre-loaded<br />

prior to installation because they are wet cells<br />

and could not be carried horizontally in the<br />

aircraft. The central electronics tube was preloaded<br />

however, and was placed in operation prior<br />

to leaving the <strong>AIDJEX</strong> main camp. The electronics<br />

were running on their internal secondary battery<br />

bank, in some cases continuously for 3 weeks,<br />

prior to deployment. This procedure obviated the<br />

need for turn-ON and extensive checkout in the<br />

field. While the tubes were being installed the<br />

upper antenna structure was being assembled. The<br />

upper structure was then placed on the tubes and<br />

electrical connections were made between the<br />

electronics tube and the batteries and antennas.<br />

A brief check was then performed to be sure that<br />

the H.F. transmitter, NAVSAT receiver and<br />

Location Aids were operational. Installation<br />

time averaged about 3.5 hours with a 3 man<br />

installatlon team and occasional assistance from<br />

the pilots. The installation <strong>of</strong> 8 AEBs was<br />

accomplished over a period <strong>of</strong> two months with<br />

the biggest delay due to weather. Their approximate<br />

positions along with that <strong>of</strong> the <strong>AIDJEX</strong><br />

main camp are shown in Figure 3.<br />

CONCLUDIHG REMARKS<br />

Eight AEBs are now operational on the Arctic<br />

Ice Pack. These buoys are presently being used<br />

to gather data necessary to finalize a mathematical<br />

model <strong>of</strong> the ice pack. As such, they provide<br />

barometric pressure temperature and position data<br />

on a 3 hour synoptic schedule at a fraction <strong>of</strong><br />

the cost <strong>of</strong> manned camps at similar locations to<br />

gather data. The AEBs have been designed with<br />

the flexibility to easily adapt to a wide variety<br />

<strong>of</strong> other sensors and can be used wherever periodic<br />

data over a long term is needed.<br />

The design and development <strong>of</strong> the AEB system<br />

and three prototype buoys was funded by the NOAA<br />

Data Buoy Office and the fabrication <strong>of</strong> the<br />

remaining AEBs and the Central Control Station<br />

was funded by the National <strong>Science</strong> Foundation<br />

through the <strong>AIDJEX</strong> Project Office.<br />

ACKNOWLEDGEMENTS<br />

An essential feature contributing to the<br />

success <strong>of</strong> the program was the close liaison and<br />

engineelring and scientific contributions by the<br />

<strong>AIDJEX</strong> staff members during the development<br />

phases <strong>of</strong> the program. In particular a debt <strong>of</strong><br />

gratitude is owed to Pat Martin, the <strong>AIDJEX</strong><br />

technical coordinator, for his dedication in<br />

reviewing and critiquing the design and for his<br />

devotion to the difficult task <strong>of</strong> completing the<br />

installation <strong>of</strong> the 8 AEBs in the Arctic ice pack.<br />

REFERENCES<br />

ITS Propagation Studies for <strong>Polar</strong> Research<br />

Lab DTD 10123173<br />

AKIMA, H., "Modulation Studies for IGOSS",<br />

ESSA Technical Report ERL 172-ITS 110,<br />

June 1970<br />

Hatfield, D. N. and DeHass T., "System<br />

Design Considerations for the Development<br />

<strong>of</strong> the IGOSS Ocean Data HF Transmission<br />

System", ESSA Technical Report ERL 158-ITS<br />

101, March 1970 .<br />

Coachman, L., "Water Masses <strong>of</strong> the Arctic"<br />

Proceedings <strong>of</strong> the Arctic Basin Symposium,<br />

Arctic Institute <strong>of</strong> North America,,Page<br />

145, October 1962<br />

Maykut, G. and Untersteiner, N., "Numerical<br />

Prediction <strong>of</strong> the Thermodynamic Response <strong>of</strong><br />

Arctic Sea Ice to Environmental Changes",<br />

Page 54-57, Rand Corporation Memorandum<br />

RM-6093-PR. November 1969<br />

Brown, W. P., ''<strong>AIDJEX</strong> System Phase I,<br />

Final Test Report for Ice Island T3 Antenna,<br />

Battery and Structure Tests", <strong>Polar</strong> Research<br />

Lab, Inc., Report TR005, October 1974<br />

18 IEEE OCEAN '75 - !X3


Sensor Word<br />

Description - Bits<br />

TABLE I<br />

Description<br />

NAVSAT Word<br />

sync 24 Sync<br />

Buoy I.D. 4 Buoy I.D.<br />

Julian Day # 10 Julian Day<br />

Word I.D. 1 Word I.D.<br />

Word # 4 Word #<br />

Engr Data 5 Engr Data<br />

Press Sensor #I 10 Buoy,Time <strong>of</strong> Day<br />

Press Sensor 12 10 Satellite I.D.<br />

3 Meter Temp 10 Re1 ZULU Time (SAT)<br />

1 Meter Temp 10 Doppler Count 81<br />

Spare $1 10 Doppler Count 12<br />

Spare #2 - 10 Doppler Count 13<br />

Total 108 Doppler Count 64<br />

Doppler Count #5<br />

Total<br />

- Bits<br />

24<br />

4<br />

10<br />

1<br />

4<br />

5<br />

14<br />

5<br />

5<br />

24<br />

24<br />

24<br />

24<br />

54 - IEEE OCEAN '75 19


Ill --<br />

- KEY<br />

AIUJEX WIN WW<br />

AEBa 0<br />

4<br />

4<br />

VI<br />

I<br />

VI<br />

VI<br />

FIGURE 2<br />

AEB STRUCTURE<br />

FIGURE 3 AEB POSITIONS AT END OF SPRING 1975


AIR DROPPABLE RAMS (ADRAMS) BUOY<br />

W. P. Brown<br />

<strong>Polar</strong> Research Laboratory, Inc.<br />

123 Santa Barbara Street<br />

Santa Barbara, California 93101<br />

E. C. Kerut<br />

NOAA Data Buoy Office<br />

Mississippi Test Facility<br />

Bay St. Louis, Mississippi 39520<br />

Abstract<br />

The ADRAMS buoy was developed to provide<br />

remote tracking <strong>of</strong> drifting sea ice near the<br />

Arctic coast. The air droppable feature was<br />

employed to reduce the high cost <strong>of</strong> deployment<br />

inherent to manual installation and to provide<br />

access to, deployment areas and seasons <strong>of</strong> the<br />

year not previously suitable. ADRAMS contains<br />

a 401.2 MHz transmitter and suitable digital<br />

encoding to allow it to be received by the<br />

NIMBUS-6 satellite. This satellite contains a<br />

random access measurement system (RAMS) package.<br />

The RAMS system determines the position<br />

<strong>of</strong> the ADRAMS buoys to an accuracy <strong>of</strong> better<br />

than 5 €31 thru doppler measurements <strong>of</strong> the<br />

received signal. The buoy is deployed via its<br />

own parachute and is designed to survive and<br />

properly orient its antenna on any type <strong>of</strong><br />

terrain. The 80 pound package contains enough<br />

batteries for 7 to 8 months operation at surface<br />

temperatures as low as -5OOC. Although<br />

the original ADRAMS was designed for tracking<br />

only, it has been modified to incorporate a<br />

capability for sensor data telemetry. The RAMS<br />

system in the NIMBUS-6 satellite accepts 32 bits<br />

<strong>of</strong> data from each transmission. Nineteen ADRAMS<br />

buoys have been deployed thus far; 17 in the<br />

Arctic and 2 in the Antarctic. The air drops<br />

have been 100% successful.<br />

1. Introduction<br />

The Bureau <strong>of</strong> Land Management, in conjunction<br />

with the Arctic Ice Dynamics Joint Experiment<br />

(<strong>AIDJEX</strong>) program, developed an urgent need for a<br />

device to allow remote tracking <strong>of</strong> the Arctic<br />

ice pack drifting near shore. This need resulted<br />

from the expansion <strong>of</strong> oil activities on the<br />

northern coast <strong>of</strong> Alaska and from problems encountered<br />

during 1975 in attempting to ship<br />

large amounts <strong>of</strong> material to the oil-rich Prudhoe<br />

Bay area.<br />

Because <strong>of</strong> the urgent nature <strong>of</strong> the program,<br />

it was probable that the buoys would have to be<br />

deployed,during the all-dark period <strong>of</strong> winter<br />

in the Arctic and under adverse weather conditions.<br />

Previously-developed Arctic data buoys<br />

(~,2) were not suitable because their installation<br />

required a crew to land on the ice, an<br />

operation too hazardous for the dark <strong>of</strong> winter.<br />

21<br />

Therefore, the concept was evolved for a small<br />

buoy which could be deployed from an aircraft via<br />

parachute. With the constraint <strong>of</strong> small size,<br />

the obvious choke <strong>of</strong> a tracking scheme was the<br />

NIMBUS-6 satellite Random Access Measurement<br />

System (RAMS). This system is being used succese<br />

fully in other buoy programs. The buoy was given<br />

the acronym ADRAMS (Air Droppable RAMS). The<br />

NOAA Data Buoy Office was selected to spearhead<br />

this program because <strong>of</strong> its wide experience in<br />

developing data buoys for the open ocean as well<br />

as the Arctic.<br />

A contractor, <strong>Polar</strong> Research Laboratory, <strong>of</strong><br />

Santa Barbara, California was selected to perform<br />

the ADRAMS design, development, and fabrication.<br />

2. Design Considerations<br />

The following factors provided the major constraints<br />

on the overall system design.<br />

1.<br />

2.<br />

3.<br />

4.<br />

5.<br />

6.<br />

7.<br />

8.<br />

The fabrication <strong>of</strong> the first group <strong>of</strong><br />

buoys had to be complete within six months<br />

<strong>of</strong> the program start to allow deployment<br />

for the critical winter season. This<br />

factor constrained the use <strong>of</strong> electronic<br />

components, materials and energy sources<br />

to those that were readily available.<br />

The size <strong>of</strong> the buoy could not Cxceed the<br />

dimensions <strong>of</strong> the parachute door on available<br />

deployment aircraft.<br />

The electronic subsystems and mechanical<br />

structure had to withstand the landing<br />

impact <strong>of</strong> a parachute landing on smooth<br />

and rough ice.<br />

The parachute had to be disconnected from<br />

the buoy after landing to prevent the buoy<br />

from being dragged across the ice by high<br />

winds.<br />

The rate <strong>of</strong> change <strong>of</strong> oscillator frequency<br />

with temperature had to be minimized, wittr<br />

out using power, to optimize tracking<br />

accuracy.<br />

The system was to have a minimum life <strong>of</strong><br />

six months.<br />

The system was to withstand the<br />

surface temperature extremes <strong>of</strong><br />

-5OOC.<br />

ice pack<br />

t10 to<br />

The buoy hull had to be water t ght in the


event that a melt pond should form around<br />

the buoy during the sumncr.<br />

9. The antenna was to remain horizontal<br />

regardless <strong>of</strong> the landing attitude.<br />

10. The cost was to be minimized since the<br />

buoys must be considered expendable.<br />

Individually, none <strong>of</strong> the above factors was<br />

difficult to achieve, but taken together they<br />

presented an interesting design problem, The<br />

requirement for completed buoys in six months<br />

posed the biggest difficulty in that time did<br />

not permit testing a variety <strong>of</strong> approaches prior<br />

to freezing the design. Fortunately, the selected<br />

approach proved satisfactory; the buoys were<br />

delivered in time and deployment was 100% successful.<br />

3. System Description<br />

A sketch <strong>of</strong> the ADRAMS buoy is shown in Figure<br />

1. The buoy consists <strong>of</strong> a 22-inch diameter polycarbonate<br />

sphere mounted on a 15-inch diameter,<br />

12-inch high foam crash pad. The electronics,<br />

antenna, and battery pack form a single unit<br />

inside the sphere and are free to rotate around<br />

both vertical and hcrizontal. axes on teflon bearings.<br />

The electronics module contains a pendulus<br />

weight so that after deployment, regardless <strong>of</strong><br />

the final resting position <strong>of</strong> the sphere, the<br />

antenna will be properly oriented for optimum<br />

satellite reception.<br />

The system is powered by newly-developed<br />

inorganic lithium batteries. These batteries<br />

have extremely high energy density on a weight<br />

and volumetric basis and allow operation down to<br />

the -5OOC temperature limit <strong>of</strong> the system. The<br />

battery pack weighs less than 9 pounds and provides<br />

an expected life <strong>of</strong> 7 to 8 months.<br />

A special ruggedized BTT (Buoy Transmit Terminal)<br />

was developed to survive the shock <strong>of</strong> an<br />

air drop as well, as the low temperature extremes<br />

<strong>of</strong> the Arctic ice pack. The electronics were<br />

bui1.t in modular form to facilitate expansion and<br />

-to allow adaption to other form factors. In<br />

addition, a "modified canted turnstile" antenna<br />

was developed from a design for the SYS satellite<br />

(3). This antenna provides circular polarization<br />

and has a low vertical pr<strong>of</strong>ile consistent with<br />

the buoy's spherical configuration. Pictures <strong>of</strong><br />

the various elemenzs <strong>of</strong> the buoy are shown in<br />

Figure 2.<br />

In operation the buoy transmits a signal for 1<br />

second each 62 seconds with a nominal radiated<br />

power <strong>of</strong> 31 dBM. The format <strong>of</strong> this transmitted<br />

signal is shown in Figure 3. A zero phase reference<br />

signal with no modulation is sent for a<br />

nominal period <strong>of</strong> 350 msec. During this period,<br />

the NIMBUS-6 satellite acquires phase lock and<br />

establishes its phase reference. The next<br />

approximately 640 msec <strong>of</strong> the signal contains the<br />

message and is phase modulated t60° at a bit rate<br />

<strong>of</strong> 100 Hz. The first 30 bits <strong>of</strong> the message contain<br />

the bit sync and frame sync followed by the<br />

platform identification number. The next two<br />

bits are used to provide up to 4 variations on<br />

the 12 bits <strong>of</strong> data which follows.<br />

The NIMBUS-6 Satellite is in polar orbit, COT"-<br />

pleting one revolution <strong>of</strong> the earth approxinatclv<br />

once each 108 minutes. The number <strong>of</strong> orhits<br />

which are in view <strong>of</strong> a platform on the earth's<br />

surface varies with the platform latitude. i'uur<br />

to six passes a day are typical near the equator<br />

to 11 or 12 passes near the poles. The number<br />

<strong>of</strong> "hits" (received transmissions) per pass also<br />

varies, depending on the elevation angle from<br />

the platform to the satellite at the closes point<br />

<strong>of</strong> approach. The variation in hits is typically<br />

3 at low elevation angles to 15 at high elevation<br />

angles. To achieve :he stated accuracy (55 L'l)<br />

<strong>of</strong> the NIMBUS-6 tracking system, at least two <strong>of</strong><br />

three consecutive passes must be received with at<br />

least 3 hits per pass (4).<br />

4. Electronic Design<br />

A block diagram <strong>of</strong> the electronic subsystem is<br />

shown in Figure 4, The oscillator is a temperature<br />

compensated crystal osciliator (TCXO) at a<br />

frequency <strong>of</strong> 3.134375 1'Mz. A higher frequency<br />

crystal could have been used, with less postmultiplication<br />

needed, but the 3 MHz crystal is<br />

inherently more stable and easier to compensate<br />

over the required wide temperature range. 'The<br />

3 PMz oscillator signal is multiplied to 50 PIHz<br />

in a X16 multiplier and then to the final 401.2<br />

MHz carrier frequency with a X8 multiplier. The<br />

carrier signal is then phase modulated by the 3-<br />

state phase modulator. The phase modulator uses<br />

a hybrid coupler for ease <strong>of</strong> impedence matching<br />

and uses pin diode switched transmission lines.<br />

Individual 10 volt logic levels from the COSPIOS<br />

digital logic selects the phase state as Oo, 60'<br />

or -6OO. The phase modulator drives the power<br />

amp, which is a single chip hybrid circuit capable<br />

<strong>of</strong> up to 2.4 watts output. The digital<br />

encoder and timer generates the bit sync, frame<br />

sync and platform I.D. as MkVCHEZTER CODED 100 Hz<br />

data which is applied to the +60 phase modulation<br />

inputs. It also generates a 0' phase signal<br />

at the beginning <strong>of</strong> the transmit cycle for 350<br />

msec. The timer portion <strong>of</strong> the digital module<br />

contains a crystal oscillator and countdown<br />

circuit to generate the basic transmit timing<br />

cycle <strong>of</strong> one second per 62 seconds. The one<br />

second transmit gate is applied to the power<br />

switching and regulator circuit which provides 3<br />

switched regulated voltage to the multiplier and<br />

power amplifier stages.<br />

The battery supply provides continuous power<br />

to the TCXO and Digital Encoder and Timer. These<br />

two units have an average drain <strong>of</strong> 12.7 milliamps.<br />

The remaining units have an average drain <strong>of</strong> 7.3<br />

milliamps with their 1/62 duty cycle. The total<br />

average drain is 20 milliamps. Five parallel<br />

banks <strong>of</strong> 4 inorganic lithium cells in series are<br />

used as the battery supply. These cells are a<br />

relatively new development by General Telephone h<br />

Electronics Laboratories. They are hermetically<br />

sealed and therefore do not have the SO2 corrosion<br />

problem <strong>of</strong> the more common organic lithium<br />

cells. They provide very high energy on a weight<br />

and volume basis with reliable operation well<br />

below the required -5OOC. The cells weigh 6.5 oz,<br />

have a terminal voltage <strong>of</strong> 3.6 volts, a flat discharge<br />

characteristic, and a capacity sf 22 amp<br />

22


hours. The twenty-cell battery pack has a voltage<br />

<strong>of</strong> 14.4 volts and a capacity <strong>of</strong> 110 amp hours.<br />

A single additional cell is used to provide<br />

switch and regulator bias. Expected life from<br />

the above pack is 7.5 months.<br />

The antenna is a modified canted turnstile<br />

antenna which provides right hand circular polarization<br />

and good hemispherical coverage. It consists<br />

<strong>of</strong> four quarter wave vertical stubs canted<br />

45O tangentially around a half wave diameter<br />

circle. Each <strong>of</strong> the stubs is fed 90' out <strong>of</strong><br />

phase with the preceding one, progressing around<br />

the circle.<br />

Although the original ADRAMS was designed for<br />

tracking only, there was soon a requirement to<br />

add a data capability. Several units were modified<br />

to incorporate a precision barometer and an<br />

internal temperature sensor. Both sensor outputs<br />

were converted to 10-bit digital words and interfaced<br />

to the digital encoder. The temperature<br />

sensor range is +30° to -5OOC with better than<br />

. l0C resolution. The barometer pressure range<br />

is 950 to 1050 millibars with .1 millibar resolution.<br />

Since the buoy had to be water tight,<br />

a special teflon membrane was used for the<br />

pressure port which allowed barometric pressure<br />

changes to enter the buoy but kept moisture out.<br />

5. Mechanical Design<br />

The major mechanical tasks associated with<br />

the development <strong>of</strong> ADRAMS included the following:<br />

development <strong>of</strong> a subsystem for paradropping<br />

ADRAMS and releasing the parachute after ice<br />

impact; packaging the system to survive ice<br />

impact; making provision for maintaining the<br />

antenna in a vertical position regardless <strong>of</strong><br />

orientation <strong>of</strong> the overall package; and selecting<br />

materials compatible with the impact loads<br />

and sub-zero temperatures.<br />

The mechanical design consists <strong>of</strong> four subsystems:<br />

1) the inner gimbal which contains the<br />

antenna and electronics; 2) the sphere which<br />

houses the gimbal; 3) the impact pad which<br />

-ahsorbs the shock <strong>of</strong> the landing and 4) the<br />

parachute and its releasing mechanism. These<br />

are shown in Figures 1 and 2.<br />

Gimbal Subsystem<br />

In order to maintain the antenna in a vertical<br />

position, the entire electronics assembly was<br />

designed as a pendulous self-leveling gimbal<br />

revolving inside a 22" diameter sphere. Tlie<br />

gimbal apprcach was found to be much simpler and<br />

more reliable than attempting to deploy automatically<br />

a fixed antenna in an upright position<br />

on the rough surface <strong>of</strong> the ice pack. The gimbal<br />

insures that the antenna will be vertical not<br />

only upon,landing but also during any future<br />

disturbances due to wind, ice surface changes<br />

due to movement, melting or polar bear interference.<br />

The weight <strong>of</strong> the entire gimbal including the<br />

electronics, batteries and antenna is 50 pounds<br />

and rests on 4 teflon bearings which are free to<br />

slide over the inside surface <strong>of</strong> the sphere.<br />

These bearings are spring-mounted onto the gimbal<br />

allowing them to retract on impact. The deceleration<br />

forces are then picked up by a polyurethane<br />

pad which distributes the impact load<br />

evenly over the bottom <strong>of</strong> the sphere.<br />

The teflon bearings have a low coefficient <strong>of</strong><br />

friction but to further enhance the self-leveling<br />

capability a lubricant was used on the inside<br />

<strong>of</strong> the polycarbonate sphere. Several oils and<br />

greases were tested for both lubrication and<br />

freezing properties. A silicone base grease<br />

was selected which has excellent lubricant qualities<br />

to temperatures below -5OOC.<br />

Buoy Hull<br />

A number <strong>of</strong> materials and fabrication<br />

approaches were considered for the outer sphere.<br />

The antenna could not tolerate metals within<br />

its pattern, so the sphere had to be non-metallic.<br />

It also had to be capable <strong>of</strong> withstanding<br />

the impact loads. Vacuum formed acrylic and<br />

ABS plastic hemispheres were tried, but the<br />

manufacturing technique produced a very thinwalled<br />

apex (about 118") thereby reducing the<br />

strength <strong>of</strong> the sphere. Polycarbonate plastics<br />

were then investigated. These plastics have<br />

about 16 times the impact strength <strong>of</strong> ABS and<br />

40 times the impact strength <strong>of</strong> acrylics. In<br />

addition they can be "rate-formed" to provide<br />

a uniform wall thickness. Rate-forming involves<br />

heating powdered resin inside a complete spherical<br />

mold which is heated and rotated about all<br />

axes. The result is a sphere with a smooth<br />

inside s-urface. The sphere is cut in half and<br />

fiberglass flanges are installed.<br />

A teflon gasket material is used to seal the<br />

two hemispheres. Caulking was rejected because<br />

<strong>of</strong> the possibility <strong>of</strong> leaks into the sphere<br />

interferring with the gimbal, Twelve 318" x 2"<br />

nylon bolts are used as fasteners for joining<br />

the hemispheres.<br />

The flange Ltself is square-shaped instead <strong>of</strong><br />

round to reduce any tendency to roll in a wind.<br />

Shock Absorbing Pad<br />

The impact pad was designed to absorb the<br />

shock <strong>of</strong> the landing and to house the parachute<br />

release switches. The shock absorbers are cubes<br />

made <strong>of</strong> polystyrene foam that crush at a constant<br />

force level. For a rate <strong>of</strong> descent <strong>of</strong> 20 feet/<br />

sec and a desired deceleration <strong>of</strong> 20 g's, a<br />

deceleration distance <strong>of</strong> 3.7 inches is required.<br />

The polystyrene used yields at 35 psi dnd<br />

crushes to 30X <strong>of</strong> its original depth so that the<br />

proper area nf material can be determined. To<br />

provide a broad support area, the impact pad was<br />

designed as a 15" diameter cylinder with the<br />

required 16 in2 <strong>of</strong> polystyrene distributed<br />

around the perimeter in 2" x 2'' blocks. To<br />

provide lateral stability, the cylinder was<br />

divided into 3 layers <strong>of</strong> 2" cubes with each<br />

layer separated by a wood disk. This provided<br />

for a rigid structure whose crush force would<br />

produce a 20 g deceleration in a 60 pound load<br />

impacting at 20 feetlsec.<br />

Parachute-Release Subsystem<br />

The parachute is designed for a d:op rate <strong>of</strong><br />

18 to 20 feet/sec with an 80 pound payload. The<br />

23


chute supplied by Paranetics Inc. is constructed<br />

in a "paraform" shape rather than the typical<br />

hemispherical design. The paraform is a modified<br />

cross which provides a highly stable, low-oscillation<br />

descent thus minimizing impact side loads.<br />

The chute must be released immediately after<br />

initial impact to prevent being dragged by the<br />

wind. To accomplish the release, an explosive<br />

cutter is mounted on the chute's main shroud,<br />

actttated upon landing by the impact switches<br />

anda four-volt battery. Each <strong>of</strong> the four<br />

switches located in the impact pad consists <strong>of</strong><br />

a large "needle" and a metallic foil imbedded in<br />

a 2"polystyrene cube. Upon impact the needle<br />

penetrates the foil and closes the circuit.<br />

Drop Testing<br />

Several drop tests were conducted both statically<br />

from a platform and by airplane. The static<br />

drops were without parachute at a height <strong>of</strong><br />

6'2"to achieve a velocity <strong>of</strong> 20 feet/sec, which<br />

is equivalent to the actual rate <strong>of</strong> descent <strong>of</strong><br />

the parachute. These first drops were primarily<br />

to test the impact pad and cutter switches but<br />

valuable structural design information was also<br />

gained. Four local test air drops were then<br />

conducted to observe the overall mechanical<br />

operation <strong>of</strong> the buoy.<br />

6. Deployment<br />

Deployment is extremely simple. It can be<br />

handled by any aircraft having an opening <strong>of</strong><br />

25" x 40". The buoys are fully operating when<br />

they are placed aboard the aircraft. They<br />

weigh only 90 pounds including the parachute and<br />

therefore deployment can be handled by one man.<br />

The aircraft should be above 300 feet and the<br />

speed should be below 100 knots. For accurate<br />

placement <strong>of</strong> the buoy in a specified target area,<br />

a low altitude <strong>of</strong> 400 to 500 feet is preferred.<br />

The deployment sequence i s as follows:<br />

1. The release battery voltage is checked.<br />

2. The explosive cutter connector is checked<br />

for absence <strong>of</strong> power and then connected<br />

to the cutter.<br />

3. Buoy transmissions are checked with a<br />

small sonic indicator.<br />

4. The static line is connected.<br />

5. The parachute door is opened.<br />

6. The base <strong>of</strong> the buoy is set on the door<br />

sill and the system tipped out.<br />

The design <strong>of</strong> the chute is such that the buoy<br />

descends within a few degrees <strong>of</strong> vertical and<br />

lands on its crash pad. The foam crash pad compresses,<br />

absorbing the landing shock. At least<br />

one <strong>of</strong> the'switches built into the crash pad is<br />

actuated by the compression and in turn actuates<br />

a guillotine cutter which separates the chute<br />

from the buoy. The buoy falls over on its side<br />

after impact and the flanges bite into the snow<br />

crust thus stabilizing the housing against being<br />

blovn by the wind. The electronics package<br />

rotates on its teflon bearings to assume its<br />

prc 'er orientation.<br />

7. Results<br />

Thirteen <strong>of</strong> the ADRAMS buoys have been ,iir<br />

dropped onto the Arctic ice pack as <strong>of</strong> May 1976<br />

without a failure. The initial delivery <strong>of</strong> eight<br />

buoys were air dropped in December <strong>of</strong> 1975 and<br />

are now completing six months <strong>of</strong> operation. Six<br />

more ADRAMS have been surface-deployed in the<br />

Arctic and Antarctic for various applications.<br />

OnIy one <strong>of</strong> these bdoys (in the Antarctic) is<br />

operating erratically.<br />

8. Summary and Conclusions<br />

The ADRAMS development has proved the feasibility<br />

<strong>of</strong> the small air drop buoy concept for<br />

use in a hostile environment. Applications thus<br />

far have included general ice pack movement,<br />

marking <strong>of</strong> specific pieces <strong>of</strong> ice for future<br />

relocation (Ice Island T-3) and acquisition <strong>of</strong><br />

barometric pressure data in remote parts <strong>of</strong> the<br />

Arctic ice pack. Other possible applications<br />

include temperature, smoke and humidity measurements<br />

in remote inaccessible areas <strong>of</strong> forests<br />

for fire prevention, seismic and air pollution<br />

measurements in remote areas.<br />

The use <strong>of</strong> the self positioning internal<br />

gimbal package, which allows deployment in rough<br />

terrains and without choosing an ideal landing<br />

spot, limits the sensor selection to those that<br />

can be contained within the package. With further<br />

development external sensors could he<br />

accomodated usine, short range telemetry techniques.<br />

An open ocean version <strong>of</strong> ADRAMS is currently<br />

under development which is primarily designed<br />

to be thrown from the deck <strong>of</strong> a ship <strong>of</strong> opportunity.<br />

However, this buoy called COSRAMS<br />

would also be suitable for air deployment.<br />

1.<br />

2.<br />

3.<br />

4.<br />

References<br />

Brown, W. P. and Kerut E. G., "Arctic<br />

Environmental Buoy System", IEEE Ocean '75,<br />

page 50.<br />

Burke, S. P. and Buck, B. M., "The SYiU'WYS<br />

Ice Station", IEEE Ocean '75, page 413.<br />

Gregorwich, W., "A hovel VHF Turnstile<br />

Antenna for the SMS Satellite", National<br />

Telecommunications Conference 1972, page<br />

36F-1.<br />

"NIMBUS-F Random Access Measurement System<br />

(WS) Platform Interface Specification",<br />

NASA Goddard Space Flight <strong>Center</strong> Document,<br />

dated September 1973.<br />

24


,v<br />

PAFORM CHUTE ,<br />

ADRAMS<br />

1 SHROUD CUTTER<br />

ELECTRONICS<br />

RESET SwITCY ,<br />

A<br />

IMPACT P4D<br />

- __<br />

--.<br />

POLY STYRENE<br />

PADS<br />

-'L CUT TE R<br />

SWITCH<br />

Figure 1 Cutaway View<br />

25


6'"<br />

IqORGANIC LITHIUM BP.TTEI?Y<br />

IMPACT CUSHION<br />

c 1<br />

INTER1 OR ELECTFONI CVSATTERY PACKAGE<br />

1.<br />

CI RCIJLARLY POLARIZED ANTENNA<br />

Fig. 2.<br />

COMPLETE ilDRPMS


7640 6-4 mSec -1<br />

980 2 26.4 msec J<br />

FIGURE 3<br />

ADRAMS TRANSMIT FORMAT<br />

ANT.<br />

TCXO<br />

3.134375 MHz<br />

X16 X8 PHASE POiJER<br />

c v AMPLIFIER<br />

Multiplier Multiplier MODULATOR<br />

401.2 MHz<br />

U<br />

r<br />

b A 4 4<br />

-<br />

1<br />

d<br />

1.<br />

+V +vs<br />

BATTERY<br />

POWER<br />

SWITCH<br />

SUPPLY<br />

L<br />

REGULATOR<br />

PHASE<br />

YODULATION<br />

XMT. DIGITAL DATA<br />

r- ENCODER 4 SIGNAL<br />

L TIMER<br />

CONDITIONER<br />

FIGURE 4<br />

ADRAMS ELECTRONIC SUBSYSTEM BLOCK DIAGRAM<br />

U<br />

SENSORS<br />

27


OM<br />

(AD RAMS)<br />

UREMEN~<br />

\ AIRDROP<br />

N<br />

m<br />

FOAM IMPACT<br />

ENERGY AB SORB ER<br />

. , _-----<br />

ELECTRON I C S, SEN S 0 RS<br />

AND BATTERIES HOUSING


The Synrams Ice Station<br />

Samuel P. Burke and Beaumont M. Buck<br />

<strong>Polar</strong> Research Laboratory, Inc.<br />

ABSTRACT<br />

A low power, unattended, ice station for collecting<br />

data has been developed to collect synoptic<br />

environmental data in polar regions for a<br />

period <strong>of</strong> two years. An array <strong>of</strong> 10 <strong>of</strong> these ice<br />

stations was installed 250-550 nautical miles<br />

north <strong>of</strong> the Alaskan coast during the spring <strong>of</strong><br />

1975. In each station, 24 hours worth <strong>of</strong> the<br />

most recent data, made up <strong>of</strong> eight 32-bit words,<br />

are retained in memory for burst transmission to<br />

the RAMS (Random Access Measurement System)<br />

receiver in the polar orbiting NIMBUS-F satellite.<br />

Surface platform location to a CPE <strong>of</strong> about 5 km<br />

is obtained through doppler measurement <strong>of</strong> the<br />

transmitted signal. This program is part <strong>of</strong> a<br />

continuing Arctic Research in Knvironmental<br />

- Acoustics (AREA) project sponsored by the Office<br />

<strong>of</strong> Naval Research, and was performed in cooperation<br />

with the Arctic Ice Dynamics Joint Experiment<br />

(<strong>AIDJEX</strong>) to study sea ice dynamics and<br />

underwater acoustics ambient noise.<br />

<strong>of</strong> the recently-launched NIMBUS F satellite. That<br />

system was developed primarily for NCAR's<br />

(National <strong>Center</strong> for<br />

.'<br />

Atmospheric Research) Tropical<br />

Wind, Energy Con ersion and Reference Level<br />

Experiment (TWERLE) Although designed for icecovered<br />

seas, SYNRAMS is easily adaptable to<br />

open-ocean applications by employing a somewhat<br />

different buoy hull configuration. Data is measured<br />

and stored in a solid state memory every<br />

three hours, eight times per day, (at the standard<br />

"synoptic weather" times <strong>of</strong> 00002, 03002, etc).<br />

with the newest data replacing the oldest data<br />

in memory. The NIMBUS F polar orbiting satellite<br />

receives signals transmitted by the ice station<br />

during one or more <strong>of</strong> the satellite's 13 daily<br />

orbits over the Arctic region. The periodic data<br />

collection and the digital memory, which requires<br />

only one satellite pass per day for complete data<br />

retrieval, make SYNRAMS unique in that all other<br />

RAMS users employ platforms that report data only<br />

at the time the satellite is in view.<br />

The location <strong>of</strong> the SYNRAMS array installed in<br />

the spring <strong>of</strong> 1975 is shown in Figuke 1. It<br />

INTRODUCTION<br />

Although some early models <strong>of</strong> data buoys<br />

employing platform-to-ground stations (HF), to<br />

satellite (VHF), and to aircraft (MF), have been<br />

tested previously in the Arctic, the =optic<br />

- Random Access Measurement S-ystem (SYNRAEIS) described<br />

here and the Arctic Environmental Buoy<br />

(AEB) represent the first attempt by the US to<br />

use such stations on a large scale for major<br />

scientific projects. The result <strong>of</strong> development<br />

and planning hased on years <strong>of</strong> Arctic experience,<br />

the successful deployment <strong>of</strong> these remote stations<br />

during the spring <strong>of</strong> 1975 represents a<br />

significant advance in Arctic technology. The<br />

advantages <strong>of</strong> small, l<strong>of</strong>ig-life, unattended automatic<br />

data buoys for scientific data collection<br />

over the extremely expensive manned stations is<br />

obvious.<br />

The successful development <strong>of</strong> these data buoys<br />

has been accomplished with a limited budget,<br />

tight schedule, and meager installation resources.<br />

The achievement required close cooperation, imagination<br />

and foresight between the sponsoring<br />

agency, ONR, and the <strong>AIDJEX</strong> Project Office and<br />

PRL.<br />

BACKGROUND<br />

The SYWlS ice station is designed to automatically<br />

measure and record underwater ambient<br />

FlW DATA BUOY IMSTALLATIONS<br />

noise, barometric pressure, and air emperature AT END OF SPRING 1975<br />

in polar seas. It utilizes the RAMS<br />

i<br />

capability<br />

29<br />

IEEE OCEAN '75 - 413


provides a grid from which geostrophic wind can<br />

be calculated using the barometric data. Wind<br />

force and ice station location data enable modeling<br />

<strong>of</strong> the effects <strong>of</strong> wind drag on ice strain on<br />

a large scale. Since underwater acoustic ambient<br />

noise is caused primarily by ice dynamics,<br />

the simultaneous measurement <strong>of</strong> acoustic noise<br />

levels with the wind drag/ice strain will provide<br />

a better understanding <strong>of</strong> the causes and mechanisms<br />

3f noise, possibly leading to a prediction<br />

model. An air temperature measurement capability<br />

was included to determine the effects <strong>of</strong> rapid<br />

surface cooling, and resultant thermal cracking<br />

<strong>of</strong> the ice on ambient noise.<br />

SYSTEM DESCRIPTION<br />

The SYNRAMS ice station is illustrated in Figure<br />

2. The drawing shows a cross section <strong>of</strong> 3.1<br />

,-<br />

,' ICE h-<br />

SIGNAL CONDITIONING ELECTRONICS<br />

BAROMETER 8 0<br />

FIGURE 2 SYNPAMS STATION INSTALLED<br />

IN ICE<br />

ITNIAMS<br />

>IC+ .> 1<br />

L CARBON-AIR BATTERY<br />

MEMORY & rONTROL<br />

MOD., ENCODER 8 TRANSMITTER<br />

\\<br />

Ponm CAPACITOR BANK<br />

.<br />

Prim<br />

meters thick sea ice with about ;1 half a meter<br />

<strong>of</strong> snow on top. A portable gasoline-powered ice<br />

auger was used to drill a 23 cm diameter hole<br />

through the ice. A 20 cm diameter, 5.2 meter<br />

aluminum tube was then placed in the hole.<br />

The noise measurement hydrophone and preamplifieris<br />

located 30 meters below the water level<br />

tethered to this aluminum tube by a length <strong>of</strong><br />

passing link chain. A 6.1 kg teardrop lead wei$it<br />

is connected to the end <strong>of</strong> the chain. The hydrophone/preamplifier<br />

cable is tied to the chain at<br />

0.3 meter intervals, and "haired" fairing attached<br />

to each loop to reduce the effects <strong>of</strong> current<br />

induced strumming.3 This 135 kg payload causes<br />

the 5.2 meter tube to float with its top about<br />

one meter above the ice surface. At the top <strong>of</strong><br />

the tube is a circularly polarized transmitting<br />

antenna (especially designed for this application)<br />

protected by a plastic fairing.<br />

A barometer and crystal oscillator for the RAMS<br />

transmitter is contained at thc extreme lower portion<br />

<strong>of</strong> the tube. These temperature sensitive<br />

components are located here to take advantage <strong>of</strong><br />

the very stable temperature <strong>of</strong> the water directly<br />

below the ice.<br />

The RAMS module is located above the barometer<br />

and oscillator unit. The lower portion <strong>of</strong> this<br />

module contains an energy storage capacitor bank<br />

for the RAMS 1-watt transmitter. The other portion<br />

<strong>of</strong> this module contains a power oscillator, digital<br />

encoder, power amplifier and power supply conditioners.<br />

The RAMS unit transmits a one-second<br />

data burst, consisting <strong>of</strong> 64 bits, once every<br />

minute. These 64 bits contain 32 bits <strong>of</strong> data,<br />

an identification word, and various system synchronization<br />

words. Each 32-bit data group<br />

corresponds to one three-hour synoptic sample<br />

occurring within the last 24 hours. Eight such<br />

transmissions cover a one-day memory period, and<br />

then the data repeats. The reception <strong>of</strong> at least<br />

eight consecutive transmissions within the 16-<br />

minute period <strong>of</strong> a pass is necessary to recover<br />

the timing <strong>of</strong> the data samples.<br />

The memoryand control module is positioned<br />

directly above the Rn'lS. This nodule contains<br />

all sensor signal conditioners, digital timing,<br />

data memory and control circuitry. The module<br />

also contains a test panel used in checkout and<br />

system setup. Connectors at the top <strong>of</strong> this<br />

module provide connections for primary power,<br />

hydrophone preamplifier, RAMS RF output and a<br />

system test set.<br />

Ten 1.2 volt carbon-air primary batteries- are<br />

stacked above the memory and control module. They<br />

are series connected to generate the f12 volt,<br />

PO00 amp-hour primary power source to power the<br />

station for a two-year period.<br />

A special circular polarized helix antennc is<br />

mounted on top <strong>of</strong> the aluminum tube covered with<br />

a ten inch diameter polyethelene fairing.<br />

The air temperature sensor is mounted next to<br />

the antenna on a wood block. The wood serves to<br />

thermally insulate the thermistor from the aluminum<br />

tube.<br />

FUNCTIONAL DESCRIPTION<br />

The basic block diagram for the conplete SYN-<br />

RAMS station is shown in Figure 3.<br />

The basic SYNRAMS ice station functionally consists<br />

<strong>of</strong> a solid state memory providing data for<br />

the Random Access Measurement System (WIS) satellite<br />

data relay platform. The memory receives<br />

its information from a digital data formatter<br />

controlled by crystal oscillator-based tining<br />

electronics. The formatter receives data from<br />

sensor signal Conditioners. The sensors used<br />

include a barometer, air temperature sensor and<br />

a hydrophone. The primary battery bank supplies<br />

power to assorted power conditioners which provide<br />

the necessary supply voltages required by the<br />

system.<br />

Each portion <strong>of</strong> the system will be discussed,<br />

starting with the sensors and working towards the<br />

RAMS system which is the output.<br />

414 - IEEE OCEAN '75<br />

30


... .<br />

ITU 0%<br />

FIGURE 3<br />

BLOCK DlAGRAH OF SVNW ICE STATION<br />

Analog Electronics<br />

Underwater acoustic ambient noise is measured<br />

in four 113 octave bands in the overall band 3.2<br />

to 1000 Hz, with an averaging time <strong>of</strong> 40 seconds.<br />

The noise field is sampled by a PRL model 34<br />

double bender hydrophone exhibiting'a flat sensitivity<br />

from 2 to 1400 Hz. The signals are<br />

amplified by an ultra low noise FET preamplifier<br />

at the phone. These units are mounted in a<br />

faired, neutrally buoyant body. The noise performance<br />

<strong>of</strong> the amplifierfhydrophone combination<br />

is considerably below previous Arctic minimum<br />

noise measurements.<br />

One hundred feet <strong>of</strong> three-conductor cable<br />

connects the chain-tethered hydrophone to the<br />

electronics module, passing to a post amplifier<br />

through a 0.7 Hz, two-pole, high pass filter.<br />

The filter prevents extremely low frequency water<br />

current-induced self noise signals at the hydrophone<br />

from reaching the post amplifier which<br />

could otherwise result in system saturation. The<br />

post amplifier drives the parallel bank <strong>of</strong> 113<br />

octave bandwidth filters. Each filter is buffered<br />

from the post amplifier by a pass band amplifier,<br />

the gain <strong>of</strong> which sets the individual measurement<br />

windows. The 113 octave filters use<br />

three pole pairs which are stagger-tuned to give<br />

a flat pass band response. Because <strong>of</strong> the excellent<br />

stop band characteristics <strong>of</strong> these filters,<br />

the effective bandwidth is equal to the half<br />

power bandwidth.<br />

The output <strong>of</strong> each filter drives a precision<br />

rectifier and averager. The DC averager has an<br />

RC time constant <strong>of</strong> 10 seconds. The full dynamic<br />

range <strong>of</strong> this converter was measured at greater<br />

than 50 db. The outputs <strong>of</strong> each <strong>of</strong> the four<br />

'<br />

averagets are switched to an analog-to-digital<br />

converter through a COS/MOS analog multiplexer<br />

controlled by timing and sequencer logic. The<br />

AID converter was designed,to give a five bit<br />

binary count which is proportional to the logarithm<br />

<strong>of</strong> the input voltage. The quantizing increment<br />

was selected to be 1.5 db, thereby providing<br />

the 48 db measurement range. Conversion and clock<br />

inputs are derived from a crystal oscillator<br />

countdown chain. The AID has a conversion time<br />

<strong>of</strong> 27 milliseconds.<br />

After each conversion the contents <strong>of</strong> the converter<br />

are placed into a parallel-in, serial-out<br />

shift register from which data are sent to the<br />

Random Access Memory (RAM).<br />

Digital Electronics<br />

Data Formatting. The data format for one synoptic<br />

sample period is outlined in Table 1.<br />

Running<br />

Total<br />

5 bits<br />

10 bits<br />

15 bits<br />

20 bits<br />

24 bits<br />

32 bits<br />

I Refer-<br />

Data Step ence Range<br />

Size Point<br />

Word 1 5 bits 1.5db -40dbu* 48db<br />

Word 2 5 bits 1.5db -4Odbp 48db<br />

Word 3 5 bits 1.5db -4Odbp 48db<br />

Word 4 5 bits 1.5db -40dbp 48db<br />

Temp 4 bits 3O C -4OOC 48OC<br />

Baro 8 bits 0.3906mb 950mb lOhb<br />

31<br />

IEEE OCEAN '75 - 41 5


words; a four-bit air temperature word; and an<br />

eight-bit barornettic pressure word. Typical out-<br />

put data may appear in binary as 11100, 11011,<br />

00011, 10111, 1101, 10101101. In decimal this<br />

would be 28, 27, 3, 23, 13. 173, and the resulting<br />

reduced data gives:<br />

Noise word 1 = 28(1.5) - 40 = +2 db//bb<br />

Noise word 2 = 27(1.5) - 40 = C.5 db//ub<br />

Noise word 3 = 3(1.4) - 40 = -35.5 db//pb<br />

Noise word 4 = 23(1.5) - -40 = -5.5 db//pb<br />

Temp = 13 (3) = 40 = -l.O°C<br />

Pressure = 173 (.3906) + 950 = 1017.5 millibar<br />

Temperature to Di-gital Converter. A linearizedthermistor<br />

located inside the antenna fairing<br />

is used to measure the average air temperature<br />

over a range from -40 to +5OC. The thennistor<br />

is used in a voltage divider connected directly<br />

to a simple, four-bit A to D converter. The<br />

converter is set up for a temperature step size<br />

<strong>of</strong> 3OC. A four bit COS/MOS binary counter holds<br />

the count after the conversion for transfer into<br />

a portion <strong>of</strong> the 17-bit data register.<br />

Barometer Signal Conditioner. A vibrating<br />

cylinder type pressure transducer, manufactured<br />

by Hamilton Standard, is used as a barometric<br />

pressure sensor. A digital, averaging type<br />

signal conditioner is configured to provide a<br />

range <strong>of</strong> 100 millibars (mb) from 950 to 1050 mb<br />

with an averaging time <strong>of</strong> 7.5 minutes. The<br />

transducer puts out an AC signal with a frequency<br />

which is proportional to pressure centered<br />

around 5500 Hz in the region <strong>of</strong> interest.<br />

Xeferring to the system block diagram <strong>of</strong> Figure<br />

3, the signal is divided by 128 and gated into<br />

an 8 bit binary counter by a 7.5 minute gate.<br />

The gate is derived from the binary countdown<br />

chain. The gated signal generates a finite<br />

series <strong>of</strong> counter overflows the remainder <strong>of</strong><br />

which (i.e., the count) is related to the average<br />

pressure during the period. This bechnique<br />

results in a resolution <strong>of</strong> l0hb/2 = 0.39mb per<br />

count. The absolute count-pressure rel.ationship<br />

is based on actual transducer calibration. The<br />

eight bit binary counter contents are transferred<br />

to a portion <strong>of</strong> the 817 bit shift register for<br />

entry into the memory.<br />

System Tim>ng and Control. The basic SYNRAMS<br />

system timing can best be illustrated by reviewing<br />

the System Memory and Control Timing diagram<br />

<strong>of</strong> Figure 4. The upper portion <strong>of</strong> the figlire<br />

shows the eight sample periods occurring within<br />

a 24-hour period. The "synoptic 0900" data<br />

sample timing is shown in detail. At 0855:43,<br />

power is applied to the barometer and after a<br />

28-second warmup delay, the barometer output<br />

wavetrain is gated into the digital conditioner<br />

by a 464 second gate. The gate period is positioned<br />

symmetrically about the hour change.<br />

At 0903:14 (TO) the noise measurement system<br />

power is applied and the analog averagers begin<br />

to stabilize. After 42 seconds have elapsed the<br />

averager outputs are sampled, digitized and<br />

written into the 24-hour memory as illustrated<br />

in the lower portion <strong>of</strong> Figure 4. During the<br />

42-second noise measurement period, the KhYS<br />

transmitter is disabled to eliminate the effects<br />

<strong>of</strong> RF pickup on the noise measurement electronics.<br />

The 17-bit data register shifts data into the<br />

Random Access Memory (WhI) five bits at a time<br />

with the A/D converter making conversions in<br />

between the shift groups. This process continues<br />

until all 32 bits corresponding to the 0000 synoptic<br />

sample have been intered into memory. At<br />

56 seconds after T$, all barometer and analog<br />

system power is shut <strong>of</strong>f, leaving only the nicropower<br />

digital system operating. The RAM timing is<br />

set up such that data can be written into memory<br />

at the same time that it is being read out. ?tenory<br />

reading and writing are done on alternate<br />

phases <strong>of</strong> the clock signal thereby providing for<br />

a synchronous operation between the KLXS system<br />

and the other digital acquisition system.<br />

The 256-bit memory transfers a 32-bit word to<br />

the W S system once every 62 seconds on a request<br />

from the UYS system timing and control logic.<br />

Over a 496-second period, the complete memory<br />

contents will have been read out corresponding to<br />

all data for the last 26-hour period.<br />

_____<br />

WYS System<br />

The Random Access Measurexent System (%VIS) is<br />

a satellite data acquisition and positioning system<br />

developed by Texas Instruments for NUA. The<br />

primary function <strong>of</strong> the system is to support the<br />

Tropical Wind, Energy Conversion and Refert-nce<br />

Level Experiment (TIV'ERLE). a National <strong>Center</strong> for<br />

Atmospheric Research (NCAR) program. The R;\!:S<br />

system aboard t!ie NIXBLIS F satellite reccive3s data<br />

from the SYNRAYS ice stations and stores it €or<br />

readout every 108 minutes to the ground acqnisition<br />

facility located in Fairbanks, Alaska. Til+<br />

station locational coordinates are determined ;Ind<br />

raw data sorted, recorded and furnished to scientific<br />

users by the central ground data processing<br />

facility located at the GoddArd Spacc Flight Cvnter<br />

in Greenbelt, Maryland. The power oscillntdr<br />

used in SYNLWS duplicates n design developed IIV<br />

NCAR for the TWERLE ballnon package.<br />

416 - IEEE OCEAN '75<br />

32


Power Supply<br />

The Dower- supply _ _ - consists <strong>of</strong> 10 carbon-air<br />

primary batteries, a regulator and dc-dc converter.<br />

This is sufficient to power the system<br />

for two years.<br />

PRELIMINARY RESULTS<br />

The complete array <strong>of</strong> 10 SYNRAMS was installed<br />

in May-June 1975. Preliminary tests with the<br />

recently orbited NIMBUS F satellite indicate<br />

that eight <strong>of</strong> the ten are operating satisfactorily<br />

and providing continuous data.<br />

- ACKNOWLEDGEMENTS<br />

The SYNRAMS system draws together many new<br />

ideas and technologies, without which this experiment<br />

would not have been possible.<br />

The technical assistance and advice provided<br />

by Ernest Litchfield <strong>of</strong> NCAR. John DuBoise and<br />

James Coates <strong>of</strong> Texas Instruments, and Pat Martin<br />

<strong>of</strong> <strong>AIDJEX</strong>, was invaluable.<br />

The cooperation and logistical support provided<br />

for the SYNRAMS installations by the <strong>AIDJEX</strong><br />

Office <strong>of</strong> the <strong>University</strong> <strong>of</strong> <strong>Washington</strong>, and<br />

especially Mr. Pat Martin <strong>of</strong> that <strong>of</strong>fice, along<br />

with Naval Arctic Research Laboratory, is greatfully<br />

acknowledged.<br />

The authors would like to thank NASA Goddard<br />

for its cooperation and support.<br />

Without the unending dedication and help<br />

provided by Mr. George Leidig <strong>of</strong> PRZ. over the<br />

two years <strong>of</strong> preparation, this project would<br />

not have been successful.<br />

This work was supported by Arctic Programs<br />

and Environmental Acoustic Program <strong>of</strong> The<br />

Office <strong>of</strong> Naval Research as part <strong>of</strong> project AREA.<br />

REFERENCES<br />

1. Coates, J. L., "The NIMEXJS F Random Access<br />

Measurement System (RAMS)", IEEE Trans. on<br />

Geoscience Electronics, Jan 75, Vol. GE-13,<br />

No. J, page 18<br />

2. Litchfield, E. W. and Gray, M. W.. "The<br />

TWERLE Balloon-to-Satellite Data Transmitting<br />

System", IEEE Transactions on Geoscience<br />

Electronics, Jan 1975, Vol GE-13, #1 page 39<br />

3. Urick, R. J., "Flutter Noise in Suspended<br />

Hydrophones", October 1960, Paper presented<br />

at the 16th Meeting <strong>of</strong> the Acoustical<br />

Society <strong>of</strong> America, San Francisco, Calif.<br />

33<br />

IEEE OCEAN '75 - 417


PERFORMANCE OF MET/OCEAN BUOYS IN <strong>AIDJEX</strong><br />

M. G. McPhee, L. Mangum, and P. Martin<br />

<strong>AIDJEX</strong><br />

ABSTRACT<br />

. Four data buoys equipped with air pressure and temperature<br />

sensors and ocean current meters were deployed in the<br />

Beaufort Sea in November 1975 in an attempt to study air-iceocean<br />

interaction while testing a new type <strong>of</strong> data buoy.<br />

Difficulties with the hardware point to many areas where improvements<br />

could be made in the future. Two buoys produced<br />

usable floe rotation and ocean current data for 145 and 332<br />

days. These data contain evidence <strong>of</strong> inertial oscillations,<br />

propagating mesoscale disturbances, and persistent westward<br />

currents which are consistent with similar observations taken<br />

from manned research camps on floes in the Arctic.<br />

INTRODUCTION<br />

Four meteorological/oceanographic (M/O) data buoys were deployed in<br />

November 1975 at locations chosen along the 1000 m isobath in the continental<br />

shelf break north <strong>of</strong> Alaska and western Canada to augment our meager<br />

knowledge <strong>of</strong> how near-surface currents in the most intense part <strong>of</strong> the<br />

anticyclonic Pacific Gyre interact with the continental shelf regime. In<br />

addition, the boundary layer structure inferred from the measured currents<br />

could be compared with similar measurements made at manned ice stations to<br />

estimate momentum exchange between the ice and upper ocean within the shear<br />

zone.<br />

The buoys, frozen in drifting floes, were tracked by the Random Access<br />

Measurements System (RAMS) <strong>of</strong> the NASA Nimbus 6 satellite, and were equipped<br />

to measure ocean currents at 2 m and 30 m below the ice. Surface pressure<br />

and temperature measurements were also made at each buoy, which, together<br />

with data from other ice and shore stations, were needed to produce accurate<br />

estimates <strong>of</strong> the wind field in the Beaufort Sea.<br />

35


Since the addition <strong>of</strong> oceanographic sensors to these buoys was a significant<br />

new step in the development <strong>of</strong> arctic data buoys, the program was<br />

considered to be experimental, with the performance <strong>of</strong> the hardware a major<br />

part <strong>of</strong> the results.<br />

BUOY DESIGN<br />

The buoy hull is a polyethylene tube 6 p11 long and 0.3 E in diameter<br />

which is frozen into a hole drilled through the ice. The bottom <strong>of</strong> the tube<br />

is plugged and has a fitting for the attachment <strong>of</strong> the upper current meter<br />

from which the mast and lower current meter are suspended. The mast is composed<br />

<strong>of</strong> ten sections <strong>of</strong> steel tubing, each about 3.3 m long and 2.5 cm in<br />

diameter, which are joined with a tapered pin to insure the proper orientation<br />

<strong>of</strong> the mast with respect to the buoy hull. This rigid suspension permits the<br />

use <strong>of</strong> a single magnetic compass in the buoy hull to sense the azimuth <strong>of</strong> the<br />

hull/mast assembly. The design, unfortunately, also permits 180" misalignment<br />

<strong>of</strong> the current meters themselves with respect to the mast and hull. From a<br />

ccmparison <strong>of</strong> the ice motion with the 2 m and 30 m currents, we believe that<br />

the 30 m sensor <strong>of</strong> M/O 1 and both sensors <strong>of</strong> M/O 4 were installed backwards,<br />

so that 180" were added to those bearings.<br />

The electrical cable which carries the current meter data is routed up<br />

the exterior <strong>of</strong> the buoy hull and enters through the top <strong>of</strong> the tube, eliminatir,g<br />

the need for an underwater connector and making the substitution <strong>of</strong><br />

another current meter string feasible. The top <strong>of</strong> the tube is covered by a<br />

larger diameter polyethylene "top hat" about 1 m tall which encloses the radio<br />

antenna and to which a shielded air temperature sensor is attached. Inside<br />

the buoy a vertical aluminum rack contains the radio transmitter, control<br />

electronics, pressure sensor, compass, and carbon-air cell batteries sufficient<br />

for one year <strong>of</strong> operation. This load, together with the weight <strong>of</strong> the<br />

current meters, mast, and top hat, causes the buoy to float vertically with<br />

about 0.5 m <strong>of</strong> the hull above the top <strong>of</strong> the ice. The rack is keyed to the<br />

buoy hull to preserve the alignment <strong>of</strong> the compass and the oceanographic mast.<br />

36


Sensor samples were taken every three hours, initially on the synoptic<br />

weather observation schedule. An inappropriate simplifying assumption in<br />

the electrical design caused the sample times to drift at the rate <strong>of</strong> three<br />

hours in one year. Since the same clock was used to control the 10-minute<br />

interval for frequency measurement <strong>of</strong> the barometric pressure sensor, the<br />

<strong>of</strong>fset <strong>of</strong> 4 parts in lo-" caused apparent errors <strong>of</strong> about 4 millibars.<br />

Both <strong>of</strong> these effects were discovered by careful examination <strong>of</strong> the data<br />

after the buoys were deployed, and the appropriate corrections were applied<br />

The current speed was also sampled over the same 10-minute interval, but<br />

the direct relationship <strong>of</strong> sample counts to current speed renders the time<br />

errors insignificant (0.03%). Samples <strong>of</strong> compass bearing, air temperature,<br />

and current direction were instantaneous--in the case <strong>of</strong> the latter, to<br />

avoid errors caused by averaging the 0'-360" crossover.<br />

The buoy memory holds eight synoptic sets <strong>of</strong> sensor samples, the oldest<br />

being replaced after a new sample period, so that the last 24 hours <strong>of</strong><br />

data are continuously stored in memory and available for transmission. The<br />

normal format for transmission to the Nimbus spacecraft is a one-second<br />

burst every minute. It was considered desirable to transmit the entire<br />

contents <strong>of</strong> buoy memory within 8 minutes to insure reception <strong>of</strong> a full day's<br />

data within one 15-minute satellite pass. Since the quantity <strong>of</strong> data in<br />

each synoptic sample set was twice the capacity<strong>of</strong>asjngleRllMS transmission,<br />

transmissions were made every 30 seconds. To avoid confusion in the organization<br />

and processing <strong>of</strong> data by NASA, two different identification codes<br />

were used alternately so that the" appearance was that <strong>of</strong> two independent<br />

buoys whose one-minute bursts happened to be 30 seconds apart. The data<br />

were easily recombined in processing at <strong>AIDJEX</strong>.<br />

When the spacecraft was within view (12-14 times each day at high<br />

latitudes) these data were received and recorded, together with measurements<br />

<strong>of</strong> the Doppler-shifted frequency <strong>of</strong> the received transmission, and played<br />

back to a NASA tracking station, usually to Gilmore Creek near Fairbanks.<br />

NASA performed basic processing <strong>of</strong> the raw data, computed geographic positions<br />

from the Doppler data, and furnished the results to our <strong>of</strong>fice on<br />

magnetic tapes. The position data wer.e corrected, edited, and smoothed by<br />

processing developed at <strong>AIDJEX</strong> [Martin and Gillespie, 1976; Thorndike and<br />

Cheung, 19771.<br />

37


.<br />

J<br />

BUOY PERFON4ANCE<br />

A brief summary <strong>of</strong> M/O buoy performance is listed in Table 1. In conformance<br />

with <strong>AIDJEX</strong> notation we have listed time in days from the beginning<br />

<strong>of</strong> calendar 1975; e.g., day 366 is 1 January 1976. Good environmental data<br />

were collected from M/O 1 for 145 days and from M/O 4 for 332 days as shown.<br />

M/O 1, M/O 2, and M/O 4 each produced good position data for more than 300<br />

days and are believed to have exhausted their power supplies. The premature<br />

environmental data failures are believed to be due to integrated circuit failures.<br />

M/O 3 was mishandled during installation and produced only very sparse<br />

data for the next four months.<br />

Figures 1 through 4 (adapted from Thqrndike and Cheung, 1977) show<br />

drift tracks for the buoys. M/O 1 left the air on day 452, but began transmitting<br />

again about two months later. Environmental data after that time are<br />

garbled, which is unfortunate since the buoy drift in the vicinity <strong>of</strong> the<br />

Barrow Canyon is <strong>of</strong>ten anomalously swift, and it would have been particularly<br />

useful to have surface current measurements there.<br />

A sample <strong>of</strong> the barometric pressure record is shown in Figure 5. Experience<br />

with these sensors (Hamilton Standard ''Vibrasense") has shown them to<br />

be very stable with time. The coarse resolution, about 1.4 millibar, is due<br />

to an erroneous simplifying assumption in the electrical design. It should<br />

have been 0.1 millibar. The data have been enhanced by smoothing and used in<br />

pressure analyses. It has not been possible to decode any coherent air temperature<br />

data, and a shortcoming in design is suspected.<br />

All current directions are referenced to the buoy hull, and the azimuth<br />

<strong>of</strong> the buoy hull is measured with a magnetic compass (Digicourse Mode1 215).<br />

Figures 6 and 7 show the magnetic azimuth data obtained from the M/O buoys.<br />

Early work on the use <strong>of</strong> magnetic compasses for data buoys in the Arctic Ocean<br />

suggested that weak horizontal components <strong>of</strong> the magnetic field and magnetic<br />

disturbances would be expected to cause errors <strong>of</strong> 5"-1.0" [Haugen and Dozier,<br />

19751. The raw data shown in the figures require a correction, unique to each<br />

compass and dependent on the compass direction and the local horizontal field<br />

strength, to get the actual magnetic azimuth used to define true current directions.<br />

Calibration curves taken in Seattle and checked in Barrow, where field<br />

38


strengths are known, were used with published values for field strength and<br />

declination. Figure 8 is the correction curve for the compass in M/O 4.<br />

These errors are due to the electromagnetic properties <strong>of</strong> other buoy components<br />

and could be eliminated or reduced with additional effort.<br />

The changes in corrections for declination, compass direction and<br />

field strength are small (< 1 degree). Since the buoy hull is frozen to<br />

the ice, changes in the uncorrected magnetic azimuth represent floe rotation<br />

and compass sampling errors. The resolution <strong>of</strong> the compass is about lo, and<br />

frequent changes in readings <strong>of</strong> about this size for M/O 4 suggest that the<br />

compass responded adequately to very small changes in azimuth. The lack <strong>of</strong><br />

noise in the raw magnetic bearings for long periods <strong>of</strong> time when the ice<br />

was stationary further suggests that the magnetic field was well behaved to<br />

the order <strong>of</strong> 1' or so. The winter <strong>of</strong> 1975-76 was remarkable for significant<br />

periods <strong>of</strong> no ice movement [Pritchard, 19761. The correlation <strong>of</strong><br />

compass bearing changes with significant events <strong>of</strong> ice drift, and the similarity<br />

between M/O buoy azimuths and the celestial and NavSatmeasurements<br />

<strong>of</strong> azimuths taken at the <strong>AIDJEX</strong> manned camps suggests that the magnetic<br />

bearings are accurate records <strong>of</strong> floe rotation [Thorndike and Cheung, 19771.<br />

The principal features <strong>of</strong> the floe rotation measured by magnetic compass on<br />

M/O buoys are, then these: the restrained, step-like behavior in the winter<br />

and spring in contrast to the more nearly continuous rotation <strong>of</strong> the summer;<br />

and the regular clockwise sense <strong>of</strong> the rotation in agreement with other<br />

measurements <strong>of</strong> floe rotation in the Beaufort Gyre [Rothrock, 19751.<br />

Raw data from the current meter sensors (Hydroproducts Savonius rotor<br />

and vane) are received as integer counts and converted to dimensional units<br />

using calibration data supplied by the designer. Figures 9 through 12 show<br />

calibrated data samples (8 per day) as they were received. The current<br />

bearing shown is the apparent direction <strong>of</strong> the current relative to the buoy<br />

azimuth. The scatter is large but not unexpected, particularly since no<br />

provision for vector averaging was made. From previous work under ice we<br />

expect turbulent eddies with time scales <strong>of</strong> from 5-10 minutes, and these<br />

would introduce large variations in directions sampled instantaneously even<br />

with steady drift. This would be especially true for the 2 m measurements.<br />

39


Another source for large variations on scales <strong>of</strong> a few hours is inertial<br />

oscillation <strong>of</strong> the ice cover and upper ocean. We found at the manned stations<br />

that it was not uncommon for the apparent direction at 30 m to swing full<br />

circle in one inertial period. Thus, the extreme scatter exhibited in Figures<br />

11 and 12 for the 30 m direction is expected. It does not show up in the 2 m<br />

direction because the water at that level is oscillating in phase with the ice.<br />

An interesting aspect <strong>of</strong> these oscillations is that their onset is apprently<br />

about two months earlier than was observed at the manned camps the previous<br />

summer. Presumably, the oscillations are damped when the ice is thick, but<br />

occur freely when the ice can no longer support internal stress gradients.<br />

For useful results it was clear that some sort <strong>of</strong> filtering <strong>of</strong> the current<br />

data was required, and as a first attempt we applied a ''cosine bell''<br />

running mean; i.e., each smoothed sample was calculated by averaging the corresponding<br />

unsmoothed sample with the 12 preceding and succeeding samples, all<br />

with the proper cosine weighting. '?he effect in the frequency domain is a<br />

low-pass filter with little energy content at periods shorter than 12 hours.<br />

The filter attenuates most <strong>of</strong> the energy at the inertial period, which is<br />

12.6 hours at 72"N.<br />

The filter was applied to the zonal and meridional velocity components.<br />

These were obtained by subtracting the corrected magnetic heading from the<br />

current direction, then adding the magnetic declination at the buoys' posit<br />

ions.<br />

Results <strong>of</strong> the calculations described above are shown in Figures 13<br />

through 16. We have reconverted the smooth components to speed and bearing<br />

and have shown them compared with the ice speed and bearing as determined from<br />

the smoothed satellite data. The reference frame is chosen such that the<br />

actual current at either level is obtained from the vector addition <strong>of</strong> the ice<br />

velocity and the measured current. In other words, if the water at 30 m were<br />

still, the 30 m current would, (ideally) have the same speed as the ice and its<br />

? <<br />

bearing would be 180" out <strong>of</strong> p%&e with that <strong>of</strong> the ice,<br />

With this in mind, the speed plots show many <strong>of</strong> the characteristics we<br />

have seen at the manned camp; i.e., the 30 m speed is usually close to and<br />

shows many <strong>of</strong> the same fluctuations as the ice speed. The 2 m speed also follows,<br />

but at a reduced magnitude, indicating that the water at 2 m is following<br />

40


the ice (causing reduced shear). An interesting event is apparent beginning<br />

about day 360 in the 30 m speed at M/O 1 (Figure 13). Note that the current<br />

speed is sustained at a level appreciably higher than the ice speed for several<br />

days. A similar event occurs at buoy M/O 4 about 10 days later. It is<br />

possible to conjecture that the events are from the same baroclinic disturbance<br />

which is propagating eastward at about 40 cm sec-l (the buoys are<br />

approximately 400 km apart). There is also a sustained current during<br />

February, March, and April 1976 at M/O 4 in the absence <strong>of</strong> much ice drift.<br />

The current is apparently westward, as would be expected in the southern<br />

part <strong>of</strong> the gyre.<br />

CONCLUSIONS<br />

Ocean current measurements from the buoys are encouraging for more<br />

widespread studies <strong>of</strong> near-surface currents under sea ice. Although the<br />

buoys were experimental, and suffered some shortcomings due to lack <strong>of</strong><br />

experience, they did provide new observations <strong>of</strong> inertial oscillations,<br />

propagating mesoscale features, and persistent westerly currents in the<br />

southern Beaufort Sea. The magnetic compass appears to have produced good<br />

azimuth data. We have had trouble interpreting the current meter directional<br />

measurements and consider some <strong>of</strong> the directional data suspect; however,<br />

this may be more due to prejudice from previous ice station experience<br />

rather than the actual evidence from the buoys. We do point out that when<br />

a towing velocity is provided by the ice, the directional measurement<br />

requires higher precision to determine the actual current direction to the<br />

same accuracy than if the current meter were fixed. This is something that<br />

should be considered in the design <strong>of</strong> future buoys. It also seems well<br />

within present technical capability to provide vectoT averaging <strong>of</strong> measured<br />

currents. The installation would be greatly simplified if current meters<br />

suspended more than a €ew meters below the buoy hull were equipped with compasses<br />

so that the rigid mast could be eliminated. Sensitive water<br />

temperature sensors seem feasible and would be useful, particularly during<br />

the summer months, to indicate the amount <strong>of</strong> stratification near the surface.<br />

41


AC KNO WL EDGME NTS<br />

The buoys were built by the Applied Physics Laboratory, <strong>University</strong> <strong>of</strong><br />

<strong>Washington</strong>, under contract for NOAA Environmental Research Laboratory, monitored<br />

by the NOAA Data Buoy Office and <strong>AIDJEX</strong>. Dean Haugen, Fred Karig, and<br />

Mike Dozier <strong>of</strong> APL deservetcredit for producing an entirely new type <strong>of</strong> buoy<br />

v’<br />

in an extremely short time. Thanks are due NASA, and especially Bill Seechuk,<br />

for not only routine operation <strong>of</strong> the Random Access Measurement System, but<br />

also for special help in keeping track <strong>of</strong> buoys day by day during field operations.<br />

Me1 Clarke and Dave Short <strong>of</strong> the <strong>AIDJEX</strong> <strong>of</strong>fice helped sort out the<br />

barometric pressure data. Thanks.<br />

REFERENCES<br />

Haugen, D. P., and K. M. Dozier. 1975. Magnetic direction reference sensors<br />

for high-latitude applications. APL-UW 7510.<br />

Martin, P., and C. R. Gillespie. 1976. Arctic odyssey: Five years <strong>of</strong> data<br />

buoys in <strong>AIDJEX</strong>. In Proceedings <strong>of</strong> the WMO/COSPAR Conference on Meteorological<br />

Observations from Space: Their Application to FGGE; and this<br />

<strong>AIDJEX</strong> <strong>Bulletin</strong>.<br />

Pritchard, R. 1976. An estimate <strong>of</strong> the strength <strong>of</strong> arctic pack ice. <strong>AIDJEX</strong><br />

<strong>Bulletin</strong>, 34, 94-113.<br />

Rothrock, D. A. 1975. The steady drift <strong>of</strong> an incompressible arcticicecover.<br />

Journal <strong>of</strong> Geophysical Research, 80(30), 387-397.<br />

Thorndike, A. S., and J. Cheung. 1977. <strong>AIDJEX</strong> measurements <strong>of</strong> sea ice motion,<br />

11 April 1975 to 14 May 1976. <strong>AIDJEX</strong> <strong>Bulletin</strong>, 35, 1-149.<br />

42


(McPhee, Mangum, and Martin)<br />

,<br />

TABLE 1<br />

MET-OCEAN<br />

BUOY PERFORMANCE<br />

MI0 Buoy RAMS Days Data Collected<br />

No. Platform Position Date Position Environmental Remarks<br />

1 141611420 71 32'N, day 306 307-452, 307-452 Env. data<br />

147 W (2 NOV 75) 519-697 poor after<br />

day 430.<br />

2 145111467 71 20'N, day 306 307-608 none Position<br />

149 W (2 Nov 75) only.<br />

3 1143/1175 73 44'N, day 309 . . .-420 none Damaged at<br />

c<br />

730 W (5 Nov 75) ins t allation,<br />

data<br />

sparse.<br />

4 124511273 71 N, day 307 308-640 308-640<br />

135 W (3 Nov 75)<br />

Note: Days are days <strong>of</strong> the "<strong>AIDJEX</strong> year":<br />

is day 366.<br />

1 Jan 75 is day 1; 1 Jan 76<br />

43


M<br />

75'N<br />

Station 14 e Veasurements<br />

by WYS buoy<br />

R 1273 from 308 to<br />

540. This met-ocean<br />

station was distinpuished<br />

from other<br />

XkYS buoys by having<br />

some oceanographic<br />

sensors on it and by<br />

havinp, two RAMS<br />

addresses. Data from<br />

the second platform<br />

(K 1245) was not as<br />

good as,Trom R 1273<br />

so it was ignored.<br />

Failed on 671, last<br />

good data on day 640.<br />

15OoW<br />

USA<br />

70"N<br />

Fig. 1


70°N 75'N n<br />

3<br />

P<br />

17OOW (D<br />

Y<br />

170°W<br />

Station 15. Measurements<br />

by met-ocean RAMS<br />

buoy 9 1420 from 307 to<br />

452. Data from R 1416<br />

at the same site was<br />

not used. The buoy<br />

was revived around<br />

day 520 and failed on<br />

day 697.<br />

150°W<br />

16OOW<br />

70'N<br />

Fig. 2.


70"N 75'N<br />

160°W<br />

150"w<br />

Station 16. Measurements<br />

by met-ocean<br />

160"W<br />

WMS buoy R 1467<br />

from 307 to 608.<br />

Data from R 1451 at<br />

the same site was<br />

not used.<br />

140'W<br />

150"W<br />

70"N<br />

Fig. 3.


12OOW<br />

*<br />

U<br />

Station 25. Xeasurements<br />

by met-ocean<br />

RAW buoy R 1143 .from<br />

381 to 420. Very poor<br />

data. The data from<br />

ri 1175 at the same<br />

site was also very<br />

Door.<br />

140"W<br />

llOoW<br />

13Oo1J<br />

70"N<br />

Fig. 4.


(McPhee, Mangum, and Martin)<br />

.. .<br />

0.<br />

IO<br />

IO<br />

.... e<br />

1<br />

ea.<br />

I I 1 I 1 I 1 I 1<br />

53 0 532 534 536 538<br />

1<br />

<strong>AIDJEX</strong> DAYS<br />

IO<br />

Fig. 5. Pressure measured by buoy 1245, days 530-540.<br />

48


(McPhee , Mangum, and Martin)<br />

2701<br />

261<br />

I<br />

-<br />

BUOY 1<br />

I<br />

253 -<br />

1<br />

244 -<br />

236-<br />

.I<br />

227<br />

-<br />

219 -<br />

COMPASS BEARING VS TIME<br />

26<br />

-<br />

9-<br />

-9<br />

-<br />

-26<br />

.<br />

-43<br />

'<br />

3io 3i9 3i9 338 3i7 356 366 3i5 384 393 403 4i2 4il<br />

49


(McPhee, Mangum, and Martin)<br />

BUOY 4<br />

I 1<br />

60<br />

43<br />

26<br />

9<br />

-9<br />

-26<br />

-43<br />

-6<br />

21 430 439 449 458 467 476 486 495 504 513 523 532 541<br />

COMPRSS BEARING VS TIME<br />

I<br />

I<br />

BUOY 4<br />

43<br />

-26<br />

-43<br />

-<br />

I<br />

50


(McPhee, Mangum, and Martin)<br />

a<br />

0<br />

a<br />

U<br />

0<br />

0<br />

ooooo<br />

0<br />

0<br />

0<br />

0<br />

0O0<br />

0<br />

0<br />

0<br />

-20<br />

0<br />

U<br />

O<br />

0<br />

0 SEATT.LE n<br />

U<br />

0 BARROW 0 0<br />

- 25; ' ' ' ' ' ' ' ' I<br />

-<br />

,I"','<br />

90 I80 270 360<br />

HEADING (degrees)<br />

Fig. 8. Correction curves for compass unit on M/O 4, as<br />

determined at Seattle and Barrow. The error indicates<br />

difference between compass reading and actual magnetic<br />

heading.<br />

51


(McPhee, Mangum, and Martin)<br />

""i a 25 ,<br />

I<br />

. . .<br />

.. .C<br />

I<br />

5!t .<br />

I . .<br />

I .a L. .I.<br />

I<br />

1 . 1 I .I -1 - I 1 I IJ<br />

01<br />

301 310 319 329 338 347 356 366 375 304 393 403 412 421<br />

WFRN FiFNFiflR nRTfl (?Ill VR TIPIF<br />

BUOY 1<br />

43


"i<br />

BUOY 4<br />

I<br />

43<br />

360-<br />

309-<br />

257<br />

206<br />

154<br />

103<br />

51<br />

0<br />

.. . -. . ...-..<br />

.<br />

. *<br />

.<br />

4<br />

.*. .<br />

.<br />

I<br />

*'<br />

..<br />

0 . . ,*. .; -<br />

*.<br />

... *<br />

* t .%" 'I.:<br />

e: ..<br />

- . * t<br />

.; -: *<br />

2.<br />

'2. ..: .-<br />

';<br />

...#<br />

*..<br />

** a<br />

.I * . ". ,<br />

-<br />

$8.<br />

':[ * ;<br />

' .'.. . ., *<br />

... .'. ..> .- .. ..<br />

*. *<br />

..<br />

* .-<br />

-<br />

. .-<br />

...<br />

.... .... ..<br />

I. :. ..<br />

* I<br />

* . .<br />

* * 0<br />

3<br />

:. *. 0<br />

- e.<br />

..<br />

' * . *<br />

- ;<br />

*q,-,"{. ......<br />

. 2 .. .<br />

*<br />

. ...... ."'<br />

*, * .,. :*; . *<br />

.A:-. ....<br />

:L<br />

I I# . ,.*:< *+<br />

. :. 2<br />

0:. *' , ..... . '.<br />

I 1 (rl); I 1 .... ,<br />

.) .<br />

e .<br />

a .<br />

Fig. 10. Current meter data from buoy M/O 4, 28 Oct 1975-25 February 1976.<br />

53


.*<br />

..<br />

(McPhee, Mangum, and Martin)<br />

I .<br />

GO,<br />

BUOY 4<br />

43<br />

3601 309<br />

257 1<br />

LL<br />

: 154 -<br />

%<br />

a *<br />

r e F<br />

-:<br />

.. !,.--.<br />

- --<br />

-<br />

I<br />

.I<br />

:r<br />

51-. r Y -<br />

- -<br />

P<br />

. 0.


a30gl<br />

251<br />

I<br />

60<br />

511<br />

r<br />

43<br />

BUOY 4<br />

I


(McPhee, Mangum, and Martin)<br />

sa<br />

47<br />

33<br />

25<br />

17<br />

8<br />

a<br />

50<br />

42<br />

33<br />

I<br />

N 25<br />

17<br />

8<br />

0<br />

50<br />

42<br />

33<br />

r<br />

25<br />

17<br />

u<br />

BUOY 1<br />

O<br />

301 310 319 329 338 341 356 366 315 304 393 403 412 421<br />

I'lUCCSSCD SI'CED VS TIMC<br />

BUOY 1<br />

PRnGFSSfD RFnRINO VS TIRE<br />

Fig. 13. Smoothed and corrected data, buoy M/Q 1. Speed is cm/sec,<br />

bearing is degrees clockwise from true north. Current data are<br />

apparent speed and direction relative to the drifting buoy.<br />

56


n 4,<br />

(McPhee, Mangum, and Martin)<br />

8UUY 4<br />

421 t<br />

50 r 1<br />

33<br />

4<br />

50<br />

42<br />

33<br />

S<br />

a 25<br />

17<br />

8<br />

0<br />

301 310 319 329 338 347 356 366 375<br />

tmnrccecn cwxn tic Tiwe<br />

384 393 403 412 421<br />

Fig. 14. Smoothed and corrected data, buoy M/O 4, day 301-421.<br />

57


(McPhee, Mangum, and Martin)<br />

50<br />

4 4<br />

33 I-<br />

-<br />

17<br />

at<br />

1<br />

RllrIY 4<br />

0 L -1<br />

r<br />

.I<br />

1<br />

17<br />

0<br />

a<br />

421 430' 439 449 450 467 416 486 495<br />

PHOCESStU SPEW VS TIME<br />

504 513 523 532<br />

.<br />

541<br />

BUOY 4<br />

421 430 439 449 450 467 476 406 495 504 513 523 532 541<br />

PROCESSED BEfiRlNlt VS TlflE<br />

Fig. 15. Smoothed and corrected data, buoy M/O 4, day 421-541.<br />

58


(McPhee, Mangum, and Martin)<br />

33<br />

W<br />

c.<br />

u 25<br />

17<br />

8<br />

0<br />

42<br />

33<br />

aI<br />

N 25<br />

, I<br />

BUOY 4<br />

,<br />

1<br />

i<br />

1<br />

E<br />

3U -<br />

42<br />

33<br />

-<br />

-<br />

I I 1,<br />

Fig. 16. Smoothed and corrected data, buoy M/O 4, day 541-661.<br />

59


A TEST OF BAROMETRIC<br />

PRESSURE AND TEMPERATURE MEASUREMENTS<br />

FROM ABRAMS BUOYS<br />

Pat<br />

Martin and Me1 Clarke<br />

<strong>AIDJEX</strong><br />

ABSTRACT<br />

Two buoys were tested for four months in 1977 at Barrow, Alaska,<br />

to resolve questions important to the measurement <strong>of</strong> barometric<br />

pressure and temperature from surface buoys in polar regions.<br />

The internal temperature stability was improved by insulating<br />

the buoy hull. Temperature effects on the barometric sensor and<br />

on the buoy clock were investigated. The barometers were free<br />

from drift over the four-month period. An air temperature sensor<br />

was found to be sensitive to radiation effects except when aspirated<br />

by the wind. The barometric pressure measurements were<br />

found to have some sensitivity to wind.<br />

INTRODUCTION<br />

In March 1977, two ADRAMS buoys fitted with pressure sensors and both<br />

internal and external temperature sensors were deployed on the tundra at Barrow,<br />

Alaska. Data were taken for about four months and compared with routine.hourly<br />

measurements taken by the National Weather Service (NWS), also at Barrow, only<br />

a few kilometers away. Figure 1 shows the comparison. Since satellite passes,<br />

and therefore buoy measurements, did not occur on the hour, linear interpolation<br />

was used between the hourly NWS samples and buoy measurements. All<br />

barometric pressure data were reduced to sea level according to measured<br />

elevation. Previous experience with the NWS barometric pressure measurements<br />

at Barrow has shown the errors to be 0.2 mb or less.<br />

The four-month test was undertaken to understand better the phenomena<br />

affecting barometric pressure and temperature measurements from ADRAMS (Air<br />

Droppable Random Access Measurement System) buoys [Brown and Kerut, 19761, so<br />

that uniformly good measurements could be obtained in the future without<br />

resorting to field calibrations. Specific areas <strong>of</strong> difficulty which had to<br />

61


.<br />

'-,..I..<br />

.<br />

B A R R O W W I N D S P E E D<br />

c<br />

. . .<br />

";:,<br />

..<br />

......<br />

......<br />

. .<br />

.<br />

.<br />

.<br />

. . .<br />

P R E S S U R E D I F F E R E N C E ( B A R R O W - B u o y 1 6 6 3 )<br />

.40[ .: , ', '? :, ;<br />

! ? o<br />

-.401<br />

.&O<br />

lor<br />

' 1<br />

,' . , I'<br />

......... ... ,.. . .'. .<br />

.:<br />

I ./.<br />

...<br />

. . . . . . . . . . , .,.: .: ~ .' '. .I..<br />

. . . . . '. .<br />

P R E S S U R E D I F F E R E N C E ( B A R R O W - B u o y 1 6 7 1 )<br />

T E M P E R A T U R E D I F F E R E N C E ( B A R R O W - 1 6 6 3 I N T . T E P P . )<br />

. 'I<br />

. .<br />

. .,..<br />

.<br />

. I.,. :. , ,. .: ............... ; .I... '"L.*,,-.. __<br />

'<br />

. ': ,<br />

. ;+ ._i.". ..... :,. ~. .;, I ,<br />

. .<br />

L.<br />

. . . . ..... ..,, . , . . .<br />

? '<br />

I .<br />

. .<br />

T E M P E R A T U R E D I F F E R E N C E ( B A R R O W - 1 6 6 3 E X T . T E ~ P . )<br />

, . ,. .<br />

. :. .,<br />

- I L ~ l I ~ 1 I I I I I I I I 1 I I I I I - 1 1 l - U - L - - - U<br />

82 84 86 88 90 92 94 9fi 98 100 102 104 106 108 110 112 114 Ilfi 118 120 122 124 126 128 130 !32 134 136 138 140 i42 144 146 148 150 152<br />

DRY, 1977<br />

Fig. la. Time series <strong>of</strong> pressure, temperature, and wind speed for first half <strong>of</strong> field test in<br />

1977. Pressures from buoys 1663 and 1671 are compared with NWS pressures; internal and<br />

external temperatures for buoy 1663 are compared with NWS temperatures.<br />

. .<br />

. .


~ 1030<br />

I<br />

1050 I I I I I I I I I I ( I J I ~ ~ ~ ~<br />

1040 -<br />

-<br />

= 1020<br />

1010<br />

BARROW P R E S S U R E<br />

-.BOL<br />

. ,. ~<br />

.. : . . ..I.<br />

.... 8.. .;- ....<br />

,.<br />

... . i'<br />

.......<br />

' :,. '_ ,<br />

. . . .. ..<br />

. . . '<br />

...... .<br />

. .<br />

." , . .:, ;;-,, ;,;:,. '.:" '. ' .,<br />

P R E S S U R E d l F F E R E N C E ( B A R R O M - B U 7 Y 16631<br />

m<br />

w<br />

...... .: . . . . . ...<br />

. ..... ..... . . . . . . . ":.'<br />

. . .<br />

* *., ... I r : ,' .<br />

. . . . . . ; I J-. ..I<br />

,;. ; Ir. ." . I , , 2 .' .;.....<br />

.<br />

. . . :'. :, - ,,. .;.<br />

'<br />

_ . .<br />

1 i:; . ,..-.. . ..........<br />

....<br />

' I .<br />

.. . .*:<br />

. I<br />

' I . . * I * . ' '<br />

. . . . . ..<br />

........ . '* '.' ,<br />

....... '! .. . .<br />

I . a i *<br />

. . .<br />

L . ;. 'i<br />

., ..<br />

T E M P E R A T U R E<br />

..<br />

( B A R R O W - 1 6 6 3<br />

D l F F E R E N C E<br />

T E o p , ,<br />

-4 '<br />

T E M P E R A T U R E D I F F E R E N C E ( B A R R O W - 1 6 6 1 E X T . T E M P . )<br />

-8 I I I I ' I 1 I '<br />

1 1 1 1 1 1 1 1 ' 1 1 ~ ' ~ ~ 1 1 1<br />

152 154 156 158 160 162 164 166 168 170 172 174 176 178 180 182 184 186 188 190 192 194 196 198 200 202 204 206 208 210 212 214<br />

DRY, 1977<br />

Fig. lb. Times series <strong>of</strong> pressure, temperature, and wind speed for first half <strong>of</strong> test in<br />

1977. Pressures from buoys 1663 and 1671 are compared with NWS pressures; internal and<br />

external temperatures for buoy 1663 are compared with NWS temperatures.


:


T<br />

E<br />

M<br />

P<br />

0<br />

C<br />

0<br />

-1 0<br />

-20<br />

-3 0<br />

"<br />

.- I i 88<br />

DAY 1976<br />

I 89<br />

Fig. 2. Diurnal variation in internal temperature <strong>of</strong> transparent and<br />

painted buoy hulls compared with NWS-measured air temperatures.<br />

Data shown are from an experiment in 1976 at Barrow. The insulated<br />

buoy hulls used in 1977 had internal temperature amplitudes slightly<br />

smaller than the air temperature, with phase lag <strong>of</strong> several hours.<br />

Twenty-day averages throughout the test period show that the sensors were<br />

free from drift for four months (Fig. 5). The mean pressure differences between<br />

buoys and NWS data are 0.08 and 0.25 mb and are probably due to small calibration<br />

shifts after the buoys left the laboratory but before they were deployed<br />

in the field.<br />

65


I<br />

25<br />

23<br />

t h--:!li[ NCI PiHCEl4l<br />

22<br />

20<br />

18<br />

17<br />

15<br />

13<br />

12<br />

10<br />

8<br />

7<br />

5<br />

3<br />

2<br />

-1.03<br />

.-:I=<br />

- .63<br />

- .2ir<br />

0<br />

.?O .so 1 .oo<br />

DCLP (.2403..14971 580 VWKS<br />

3E<br />

34<br />

3?<br />

30<br />

PB<br />

eG<br />

24<br />

22<br />

70<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

c<br />

4<br />

2<br />

c<br />

J<br />

36<br />

34<br />

32<br />

30<br />

28<br />

26<br />

24<br />

22<br />

73<br />

18<br />

1G<br />

14<br />

12<br />

10<br />

e<br />

0<br />

.2U .GO 1 .Od<br />

6<br />

4<br />

2<br />

Fig. 3. Pressure differences for the first half <strong>of</strong> the test. Yumbers<br />

following DELP are the mean, standard deviation, and number <strong>of</strong><br />

samples for the data set.<br />

66


f RCGXNCY<br />

13Ur - ---<br />

I<br />

120.<br />

110-<br />

YtRCfNT<br />

103<br />

90.<br />

eo -<br />

70.<br />

tu '<br />

50<br />

40 -<br />

30.<br />

2@<br />

i<br />

~<br />

10-<br />

~"l-'-LI--L.t<br />

-1 .oo - .CU 7.25<br />

PERCFCEEI<br />

tI8 BU3Y 1553 - UUOY 1c371 t10 BUOY It63 - M:J?f 1671 CIOCt( WLO COI!STAST<br />

II<br />

38<br />

36<br />

33<br />

31<br />

za<br />

26<br />

23<br />

21<br />

10<br />

15<br />

13<br />

10<br />

0<br />

5<br />

3<br />

0<br />

0<br />

ELF I .0102. .Oo'?QIJl<br />

391 VRI.UES<br />

Fig. 4. Pressure differences for the second half <strong>of</strong> the test. Numbers<br />

following DEL$ are the mean, standard deviation, and number <strong>of</strong><br />

samples for the data set.<br />

67


.5<br />

.4<br />

.3<br />

0<br />

A<br />

R 2<br />

R<br />

0<br />

W J<br />

- 0<br />

B<br />

U - .I<br />

0<br />

Y<br />

I4<br />

B<br />

- .2<br />

+- 3<br />

+<br />

8<br />

- .4<br />

-.5<br />

82 102 122 142 162 182 2 02 222<br />

DAY, 1977<br />

Fig. 5. Twenty-day averages <strong>of</strong> pressure differences between<br />

NWS ("Barrowtt) data and the two buoys (NWS - buoy). The<br />

figure shows that the barometers were free from drift for<br />

the four-month test period.<br />

Sensor Temperature Coefficients<br />

Knowing how sensitive a barometric pressure sensor is to temperature is<br />

critical to achieving good pressure measurements from ADRAMS. The temperature<br />

dependence <strong>of</strong> sensors used in this test (Paroscientific Digiquartz Model<br />

215A) is a complex function <strong>of</strong> temperature effects on several <strong>of</strong> the sensor<br />

components, principally the bellows and quartz beam. Coefficients are<br />

generally within a factor <strong>of</strong> 3 <strong>of</strong> 0.04 mb per degree Celsius, as measured<br />

at the factory over a wide range <strong>of</strong> temperatures and pressures. Barometer<br />

calibration coefficients were calculated at (and temperature corrections<br />

therefore normalized about) -26OC, or -15'F, to minimize pressure-dependent<br />

68


temperature effects in the primary operating temperature range (Fig. 6).<br />

single temperature correction curve taken at the standard pressure will thus<br />

have pressure-dependent errors <strong>of</strong> less than 0.1 mb throughout the temperature<br />

and pressure ranges <strong>of</strong> interest.<br />

There were no significant temperature-related errors when buoy pressures<br />

were corrected according to the laboratory calibrations and compared with NWS<br />

pressures over the test period (Fig. 7). This suggests not only that the<br />

laboratory calibrations were appropriate to the field application, but also<br />

that the temperature sensitivity <strong>of</strong> the pressure sensors was stable with time<br />

The latter point has been verified by repeating laboratory calibrations after<br />

field use.<br />

An important part <strong>of</strong> compensation for temperature effects is the accuracy<br />

and stability <strong>of</strong> the internal temperature sensors. Field checks <strong>of</strong> laboratory<br />

calibrations <strong>of</strong> the thermistors have agreed within l0C, which is sufficient.<br />

The lack <strong>of</strong> temperature-related errors throughout the test period further<br />

suggests that the temperature sensors are stable with time.<br />

Since temperature dependence is a function <strong>of</strong> full-scale sensor range,<br />

it is possible to reduce temperature sensitivity per unit pressure at the<br />

expense <strong>of</strong> decreased pressure accuracy by selecting a sensor with increased<br />

full-scale range. For example, a sensor with two bars full-scale rating would<br />

be expected to have half the temperature sensitivity per n%ZZibar <strong>of</strong> a one-bar<br />

sensor, but twice the aging and sensitivity to clock errors.<br />

1<br />

A<br />

C1 ock Correction<br />

To measure barometric pressure using a sensor with a frequency output<br />

requires an accurate time base from which to reference the counting <strong>of</strong> sensor<br />

output. Time base errors <strong>of</strong> one part in lo5 contribute 0.1 mb pressure errors<br />

for the sensors used here. Nonstandard initial values, long-term aging, and<br />

seasonal and diurnal temperature effects can easily combine to produce clockrelated<br />

pressure errors <strong>of</strong> the order <strong>of</strong> 1 mb. It is possible to observe the<br />

behavior <strong>of</strong> the clock through the transmissions to the satellite and correct<br />

the pressure measurements accordingly (see Appendix A).<br />

69


9<br />

1.5<br />

1.0<br />

c<br />

8<br />

w .5<br />

R<br />

930 $<br />

565<br />

1000<br />

1035<br />

1OG 0<br />

1-4 0<br />

B<br />

-. 5<br />

-1.0<br />

930 .+<br />

96'5 +<br />

I I I - 1 1 -<br />

,<br />

-68 -LO -20 0 20 40 60 80<br />

TEMPERArURE O F<br />

Buoy 1671<br />

1.0<br />

c<br />

0<br />

R<br />

R -5<br />

B<br />

E<br />

0<br />

0<br />

0<br />

0<br />

0<br />

-1.0<br />

TEMPERATURE "F<br />

Fig. 6, Temperature coefficients for the two barometers in each<br />

buoy, normalized about -15OF (-26OC) and 1000 mb: (+) =<br />

measured data, (0) = fitted curve. A single temperature<br />

correction curve has pressure-dependent errors <strong>of</strong> less than<br />

0.11 mb over the range from -40°F to +32OF, and between 1000<br />

and 1030 mb.


1 .oo<br />

I<br />

1<br />

.80<br />

.50 .60 t<br />

*<br />

Cg ,40-<br />

a 3 0 -<br />

>- .20-':.<br />

0<br />

3 .lom<br />

0-<br />

I<br />

-.loi5<br />

-.20-<br />

[YI<br />

E -.30-<br />

5 a -.40-<br />

*<br />

*<br />

*<br />

* .<br />

......<br />

.*<br />

...*:'. .. ...<br />

.....>


Clock corrections were applied routinely to the data and were effective<br />

in correcting for large nonstandard initial values. An attempt to make the<br />

correction algorithm responsive to diurnal clock variations incorporated<br />

featzres which caused random pressure errors <strong>of</strong> 0.15 mb and a significant<br />

Loss cf o%hem-Lse good ~ressurs xeasuremmts. The precision <strong>of</strong> buoy pressures<br />

is significantly better when the buoy clocks are assumed to be<br />

constact with time (Figs. 3 and 4).<br />

A more efficient aljprithm incorporating better resolution and some<br />

averaging could. be used easily, OH simpler checks <strong>of</strong> clock behavior could<br />

be done periodically by hand. The use <strong>of</strong> clocks with smaller temperature<br />

coefficients, insulation <strong>of</strong> the buoy, and relaxation <strong>of</strong> pressure accuracy<br />

requirements would reduce the need for frequent clock corrections. Selection<br />

<strong>of</strong> barometers with decreased pressure sensitivity per unit change in<br />

frequency output would increase the need for frequent clock correction.<br />

Wind-Induced Dynamic Pressures<br />

The Parge tails in the pressure difference histograms (Figs. 3 and 4)<br />

and thz ano?iaiy in thc pressure difference time series at day 140 (Fig. 1)<br />

are due to dynamic pressures associated with high winds.<br />

The pressure orifice<br />

<strong>of</strong> buoy 1663 faced northeast into the strongest prevailing winds occurring in<br />

the first half <strong>of</strong> the test period, while buoy 1671 faced the opposite direc-<br />

tno~. ligure 8 ~RQWS the distribution <strong>of</strong> winds for the test period: Figure 9<br />

S~QWS that the pressure errors depend on the component <strong>of</strong> wind on the pressure<br />

orifice. The effect is most pronounced during the first half <strong>of</strong> the test,<br />

when the highest on-axis winds occurred.<br />

The suggestion <strong>of</strong> lower buoy pres-<br />

sures during high on-axis winds for buoy 1671 may be due to sensitivity <strong>of</strong><br />

the raWS reference pressures to southwest winds, or to a protective shroud<br />

covering the orifice <strong>of</strong> that buoy.<br />

The effect <strong>of</strong> wind gusts on the pressure sample is reduced by the<br />

combination <strong>of</strong> a teflon filter in the pressure orifice and a one-minute<br />

integration period which produces a two-minute response time to pressure<br />

change,<br />

However, the horizontal orientation <strong>of</strong> the port makes the pressure<br />

measurement susceptible to the dynamic pressures <strong>of</strong> the average wind.<br />

occurrence <strong>of</strong> these errors is frequent enough to affect the sample variances<br />

The<br />

72


Nr3RTH<br />

I<br />

SC'JTH<br />

BRR40W KIN3 ORTE<br />

June - July<br />

- 8 0 . 9<br />

2 M/S PER 3i'J<br />

Fig. 8. Wind roses at Barrow for first and second halves <strong>of</strong> test period.<br />

Pressure orifice <strong>of</strong> buoy 1663 pointed to the northeast and was exposed<br />

to strong on-axis winds during the first half <strong>of</strong> the test period.<br />

in the first half <strong>of</strong> the test, but not the sample means. Smaller variances<br />

during the second half <strong>of</strong> the test reflect smaller on-axis winds during that<br />

period. Since these errors occur only during times <strong>of</strong> high winds, which are<br />

generally driven by large pressure gradients, they do not seriously degrade<br />

the usefulness <strong>of</strong> the barometric pressure measurements.<br />

AIR TEMPERATURE MEASUREMENTS<br />

The measurement <strong>of</strong> air temperature requires a sensor on the exterior <strong>of</strong><br />

the buoy hull, freely exposed to the atmosphere. The sensor must be rugged<br />

enough to withstand a severe direct impact with ice and snow when the buoy<br />

bounces during a parachute landing. Radiation effects must also be minimized.<br />

In the test, radiation effects were to be minimized by using a small<br />

(14 mil) thermistor with low thermal mass-to-surface-area ratio to enhance<br />

self-cooling. The thermistor was protected by a permanent plastic arch and<br />

a temporary plastic cup. The cup was designed to remain over the sensor for<br />

12 seconds after initial impact, but in fact when one <strong>of</strong> the buoys was dropped<br />

in a test deployment the cup disengaged prematurely; on a second attempt the<br />

cup, the arch, and the sensor itself all broke cleanly away from the buoy hull,<br />

73


Q)<br />

u o<br />

z<br />

z i 4- 4-<br />

Z<br />

+ +<br />

z<br />

f<br />

n<br />

9<br />

Ef .3<br />

v<br />

+:€E3<br />

21671<br />

Jr Fig. 9. Pressure errors<br />

as a function <strong>of</strong> wind<br />

component on the pressure<br />

orifice for first<br />

- .s<br />

-1s -10 -5 3 S i0 15 __ and second halves <strong>of</strong><br />

K!h3 SPEED X/S the test period.<br />

z z z z z<br />

Dynamic pressures are<br />

most pronounced for<br />

the first half and<br />

for buoy 1663.<br />

ending test deployments with negative results and making that buoy incapable<br />

<strong>of</strong> measuring external temperature for the rest <strong>of</strong>.the test. The temperature<br />

sensor on buoy 1663 (the one that was not dropped) functioned well throughout<br />

the data collection, ending speculation that the small thermistor would not<br />

survive the abrasion <strong>of</strong> blowing snow.<br />

Differences in air temperature between the buoy and the NWS data<br />

averaged less than l0C and showed no evidence <strong>of</strong> drift. Occasional sharp<br />

differences <strong>of</strong> about 5OC, especially during the spring, occurred at times<br />

<strong>of</strong> low wind speed (Fig. 1). The reduced conductive heat transfer with the<br />

air permitted radiative heating <strong>of</strong> the sensor during the day and radiative<br />

74


cooling at night. Temperatures are good to about 2OC for wind speeds greater<br />

than 4 m sec-' (Fig. 10). Cloud cover reduced radiation errors in the second<br />

half <strong>of</strong> the test period.<br />

Internal and external temperature measurement errors with respect to the<br />

NWS data for spring and summer are given in Table 1. Means and standard deviation<br />

<strong>of</strong> errors in daily averages <strong>of</strong> the external temperature measurements are<br />

small enough (about l0C) to suggest that daily average temperatures may be<br />

useful. Estimates <strong>of</strong> errors in daily averages might be weighted according to<br />

wind speeds calculated from the pressure measurements. We have not calculated<br />

such a weighting. Winter temperatures would be expected to be consistently<br />

too low due to constant radiative cooling and low wind speeds during that<br />

season. Under these conditions, it is also possible for a cold surface layer<br />

1-2 m thick to form, which would limit the value <strong>of</strong> winter temperature<br />

measurements. It is possible that daily averages <strong>of</strong> the internal buoy<br />

temperature would be useful under these and other conditions, especially if<br />

the radiation sensitivity <strong>of</strong> the buoy itself could be reduced.<br />

75


TABLE 1<br />

DIFFERENCES IN DAILY TEMPERATURE (NWS "BARROW" DATA MINUS BUOY DATA)<br />

Internal Temperature, OC<br />

External Temperature, OC<br />

Kinimum Maximum Mean Minimum Maximum Mean<br />

Buoy 2663<br />

March-May<br />

Me an 0.79 -0.52 0.30 1.92 0.05 0.99<br />

Standard<br />

Deviation 2.39 2.58 2.21 1.86 1.76 1.29<br />

June- July<br />

Me an -3.07 -3.20 -3.18 1.20 -0.37 0.65<br />

Standard<br />

Deviation 1.57 2.37 1.75 0.73 3.16 0.90<br />

,<br />

Buoy 1871<br />

March-May<br />

Me an 1.87 0.22 0.81 -- -- --<br />

Standard<br />

Deviation 5.50 2.33 2.03<br />

--<br />

-- --<br />

June-July<br />

Me an -2.42 -2.71 -2.45 -- -- --<br />

Standard<br />

Devi at ion 1.25 2.42 1.50<br />

-- -- --<br />

CONCLUSIONS<br />

The results <strong>of</strong> this test provide a basis for designing pressure and temperature<br />

sensors for ADRAMS buoys. They apply directly to the equipment and<br />

weather conditions actually tested, and may be instructive for other<br />

applications.<br />

Insulation <strong>of</strong> the buoy hull effectively reduced temperature variations<br />

in the buoy electronics and thus improves temperature compensation <strong>of</strong> the<br />

pressure sensor. This feature has since become standard for ADRAMS.<br />

The barometers were free from drift during the field experiment.<br />

Temperature compensation errors were not detectable. The clock correction<br />

76


algorithm removed large initial errors, but introduced 0.15 mb random errors<br />

and significant loss <strong>of</strong> data in an unnecessary attempt to correct for diurnal<br />

clock variations. Wind caused systematic pressure errors, but they would not<br />

be significant for most applications. With appropriate corrections, the barometric<br />

pressure is good to better than 0.1 mb, an error due largely to the<br />

sampling resolution.<br />

The external temperature sensors must be protected better than they<br />

were during the test. The suitability <strong>of</strong> daily average external and internal<br />

temperature measurements should be considered. If better instantaneous air<br />

temperatures are needed, smaller sensors or some other treatment <strong>of</strong> radiation<br />

effects will be required.<br />

In conclusion, it is possible to obtain uniformly good measurements <strong>of</strong><br />

barometric pressure and temperature without resorting to field calibrations.<br />

A summary <strong>of</strong> specific pressure accuracy requirements is given in Appendix B.<br />

ACKNOWLEDGMENT<br />

NASA operated the data collection system and provided the data from the<br />

field experiment at no charge. The Naval Arctic Research Laboratory provided<br />

field support for the test with the cooperation <strong>of</strong> the Arctic Project Office<br />

<strong>of</strong> the Outer Continental Shelf Environmental Assessment Program. The work<br />

was performed under the auspices <strong>of</strong> the <strong>Polar</strong> Research <strong>Center</strong>, <strong>University</strong> <strong>of</strong><br />

<strong>Washington</strong>, for the NOAA Data Buoy Office on contract no. 01-7-038-947.<br />

APPENDIX A<br />

The Digiquartz sensors are relatively insensitive to pressure (full-scale<br />

excursions in pressure produce only a 10% change in output), which puts a<br />

greater burden on counting accuracy. A sensor with one atmosphere full-scale<br />

range will have a change in output <strong>of</strong> only one part in lo5 for 0.1 mb pressure<br />

change. Accordingly, the time base will contribute 0.1 mb error for each part<br />

in lo5 variation from the assumed clock rate. While this clock accuracy is<br />

achieved easily with aodern laboratory equipment, it cannot be taken for granted<br />

over a temperature range <strong>of</strong> 8OoC and a time scale <strong>of</strong> one year.<br />

77


Laboratory calibration <strong>of</strong> the ADRAMS clock shows temperature sensitivity<br />

as Parge as one part in IO5 in 8OC (Fig. 11). Errors <strong>of</strong> a few tenths <strong>of</strong><br />

a nillibar could be expected due to temperature effects on the clock.<br />

ile it would be possible, in principle, to measure the temperature <strong>of</strong><br />

thn9 clock and to compensate (as is done with the pressure sensor), this<br />

would require laboratory temperature calibrations <strong>of</strong> the clock and another<br />

in situ temperature measurement, as well as additional data transmission.<br />

This approach would do nothing to correct for the aging <strong>of</strong> the time base.<br />

Fortunately, transmission <strong>of</strong> data through the satellite provides a<br />

means <strong>of</strong> measuring and correcting for the actual variations in the buoy<br />

clock, since accurate time measurements are associated with each data recep-<br />

~iow. It is necessary that the same clock be used to gate the barometer<br />

counts and to regulate the transmission intervals <strong>of</strong> the buoy, preferably<br />

without analog timing circuits involved in controlling the transmission<br />

interval. ADRATIS circuits meet these requirements e<br />

.- -20 0 20<br />

MPERATURE O C<br />

Fig. 11. Tenperature sensitivity <strong>of</strong> the clock for buoy 1663 from<br />

laboratory calibration.<br />

78


One part in lo5 <strong>of</strong> the nominal RAMS transmission interval <strong>of</strong> 60 seconds<br />

is 0.6 msec, which is the order <strong>of</strong> accuracy needed to determine the transmission<br />

interval. so that the clock can be corrected to 0.1 mb equivalent error.<br />

The 0.1 second resolution in measuring time in the RAMS flown on Nimbus<br />

requires that 10,000 seconds elapse between observations used in the solution<br />

for transmission interval to meet the accuracy goal <strong>of</strong> 0.1 mb. The interval<br />

between sequential satellite passes is about 6,600 seconds; this is sufficient<br />

to obtain! an accuracy <strong>of</strong> 1.5 parts in lo5, or about 1 msec uncertainty<br />

in the transmission interval.<br />

This is the accuracy obtained in the clock correction used in this<br />

study. The shortcoming introduced a 0.15 mb random error Pn the barometric<br />

pressure measurements. Clock correction based on measurements separated by<br />

two orbital periods would have been more appropriate to the test.<br />

The temperature dependence <strong>of</strong> the clocks in the field, when compared<br />

with the spacecraft clock, is seen to be <strong>of</strong> the same sign as the laboratory<br />

calibrations,'but about half the magnitude (Fig. 12). This difference in<br />

magnitude is probably due to temperature differences between the clock<br />

circuits and the internal temperature sensor located at the barometer. The<br />

results could also have been affected by clock aging, which would be indistinguishable<br />

from seasonally varying temperature effects. Direct measurement<br />

<strong>of</strong> (and correction for) clock behavior makes the distinction unimportant.<br />

APPENDIX B<br />

The following design criteria are stated with respect to pressure<br />

accuracy.<br />

If 1 mb errors are acceptable in pressure measurement,<br />

0 No temperature compensation need be used for barometers with coefficients<br />

less than about 0.04 mb per degree Celsius.<br />

0 An opaque but uninsulated buoy hull may be used.<br />

0 Only initial and long-term corrections for clock variations are necessary.<br />

If 0.3 mb accuracy is nscessaq,<br />

0 Barometers with coefficients greater than 0.01 mb/OC must be temperature<br />

79


3.20-<br />

-<br />

1<br />

t<br />

W<br />

o?<br />

2 3.051 ...:-.;.. ....<br />

I.<br />

(D<br />

*<br />

.......... .<br />

. - . .. e ."., ..<br />

4 3.05<br />

(D<br />

I. I --.. -..-. ." s .* *<br />

....- - ......<br />

"<br />

............ ... .<br />

..I<br />

.. ........<br />

...I<br />

......<br />

. . . . ..<br />

.I.. ...<br />

5 2.95<br />

. *<br />

I. . .. *. . - .... ...................<br />

I -. -..-<br />

......_._.<br />

.--.". ..<br />

i i<br />

-....). --<br />

* 7 2-90<br />

.. . . .............._........<br />

..a.. .<br />

2.85<br />

i<br />

-<br />

u<br />

w<br />

u,<br />

4-<br />

3.15<br />

3.10<br />

1<br />

2.80<br />

2 .PI5<br />

....<br />

........<br />

..<br />

.... . . . . .<br />

.I. ... ..... .<br />

"<br />

..... .--.. .."<br />

. ..<br />

" ... ..' . *<br />

, - s . .<br />

. . - . .<br />

, * e.<br />

1. ..<br />

. ."...<br />

..<br />

*. .,. a<br />

.....<br />

.....<br />

2.82'- I -<br />

1 -1-1<br />

-45.0 -35.0 -25.0 -15.0 -5.0 5 .o 15 .o -45.0 -.Qc- .J.l.cj -25.g -15.0 -5.0 5 .o 15 -0<br />

I NTERNFIL TEMPEKHTURE DEG C 1663 I NTER.!~!i-lL ,TEKT'ERFl: I.SE !lEG C 1671<br />

Fig. 12. Temperature sensitivity <strong>of</strong> the ADRAMS clocks respect to temperature here is <strong>of</strong> the same sense<br />

during the field test from satellite data. The as the slope with respect to frequency in Figure<br />

units <strong>of</strong> clock interval are tens <strong>of</strong> milliseconds. 11, but about half the magnitude. The difference<br />

The scatter is due to poor resolution in the routine<br />

to calculate clock rate, sampling resolutions,<br />

<strong>of</strong> the mean from the designed clock interval <strong>of</strong><br />

61.825 seconds is 5 msec, and is equivalent to a<br />

and the effect <strong>of</strong> winds. The slope <strong>of</strong> period with 1 mb pressure error if not corrected.


compensated.<br />

in insulated buoy hulls.<br />

Those with coefficients greater than O.lmb/OC must be used<br />

a Either an efficient clock routine or a clock with a low temperature<br />

coefficient (less than 1 lo-'/ C, -40+O0C) must be used with daily<br />

clock corrections.<br />

If 0.1 mb accwlacy is desired,<br />

0 Temperature compensation <strong>of</strong> barometers must be used, and buoy hulls must<br />

be insulated if temperature coefficients are greater than 0.01 mb/OC.<br />

0 An efficient clock correction routine must be used.<br />

These criteria are subjective and somewhat arbitrary, and they should be<br />

used with that in mind. It is hoped that they will provide a framework from<br />

which to consider design alternatives.<br />

REFERENCES<br />

Brown, W, P., and E. G. Kerut. 1976. Air droppable RAMS (AD'RAMS) buoy.<br />

In Ocean 76, Proceedings <strong>of</strong> the 1976 IntemationaZ Conference on<br />

Engineering in the Ocean Environment, IEEE Publ. No. 76 CHO 1118-9 OEC,<br />

sec. 14, D1-6.<br />

81


POSITION MEASUREMENTS OF <strong>AIDJEX</strong> MANNED CAMPS<br />

USING THE NAVY NAVIGATION SATELLITE SYSTEM<br />

Pat Martin, C. G. Gillespie, Alan Thomdike<br />

<strong>AIDJEX</strong><br />

D. Wells<br />

Bedford Institute <strong>of</strong> Oceaxography<br />

Dartmouth, Nova Scotia<br />

ABSTRACT<br />

The Navy'Navigation Satellite System was used during <strong>AIDJEX</strong> to make<br />

precise measurements <strong>of</strong> the positions <strong>of</strong> the four manned camps.<br />

Satellite signals were acquired under computer control to standardize<br />

and maximize data collection. Single-channel (400 MHz)<br />

translocation (relative positioning) was used to measure changes<br />

in distance between camps and a reference azimuth at each camp.<br />

Special attention was given to the frequency and time references<br />

for Doppler counting, and an efficient editing algorithm was used<br />

to ensure that translocation Doppler data matched. Statistical<br />

filtering after the experiment produced smoothed and evenly spaced<br />

estimates <strong>of</strong> position, velocity, acceleration, and aeimuth.<br />

An ensemble <strong>of</strong> 500 fixes from a period <strong>of</strong> virtually no ice motion<br />

was examined for position errors. Of these, 68% from a single<br />

station had errors <strong>of</strong> less than 75 m. For translocation between<br />

two stations separated by 100 km, 68% <strong>of</strong> the errors were less than<br />

20 my and between two receivers separated by 100 m at the same<br />

station 68% <strong>of</strong> the errors were less than 5 m.<br />

INTRODUCTION<br />

During <strong>AIDJEX</strong>, precise position measurements were needed to determine<br />

large-scale sea ice deformation. These measurements were taken from March<br />

1975 through May 1976 with the Navy Navigation Satellite System (NavSat)<br />

at four manned camps in the Beaufort Sea, which were deployed at approximately<br />

73'N, 156OW. (NavSat measurements made at unmanned buoys are<br />

described by Brown and Kerut [1975 and this <strong>Bulletin</strong>, pp. 15-20].) Following<br />

the field experiment, time sequences <strong>of</strong> the position <strong>of</strong> each camp and the<br />

distance between the camps were used as input to a statistical filter, which<br />

83


incorporated the positioning accuracy information to give smoothed time series<br />

<strong>of</strong> position, velocity, acceleration, and azimuth.<br />

This paper describes the positioning system and the changes made to<br />

improve its performance, among them the removal <strong>of</strong> timing errors by correcting<br />

Doppler counts; automatic selection, acquisition, and break-<strong>of</strong>f from satellite<br />

passes; and definition <strong>of</strong> a simplified editing procedure suitable for realtime<br />

fix processing which gives the best translocation matching possible.<br />

The positioning system at each camp contained two receivers, their antennas<br />

separated by 100 m, to provide azimuth data and receiver redundancy.<br />

REQUIREMENTS FOR POSITION ACCURACY<br />

The motion <strong>of</strong> a piece <strong>of</strong> sea ice adjusts to balance the forces exerted by<br />

the air and the ocean, the Coriolis force, inertia, and stress gradients within<br />

the ice. An objective <strong>of</strong> <strong>AIDJEX</strong> has been to study this response and to collect<br />

data against which theories could be tested.<br />

Motion is used here in several senses. Generally, the motion <strong>of</strong> a piece<br />

<strong>of</strong> ice is the time series <strong>of</strong> the positions <strong>of</strong> that piece <strong>of</strong> ice. The term<br />

is also used loosely to refer to quantities derived from one or several <strong>of</strong><br />

these time series--in particular, the ice velocity, acceleration, and deformation.<br />

These quantities figure into the balance-<strong>of</strong>-force equation: the<br />

velocity appears in the Coriolis force, the acceleration in the inertia term,<br />

and the deformation is related to the ice stress by a postulated constitutive<br />

law. The need to estimate each <strong>of</strong> these quantities for the force balance<br />

determined the sampling and accuracy criteria for the position measurement<br />

p r o g ram.<br />

On the basis <strong>of</strong> earlier work [Thorndike, 19741, it was decided that the<br />

highest frequencies <strong>of</strong> interest were about two cycles per day (approximately<br />

the inertial frequency at the latitude <strong>of</strong> the <strong>AIDJEX</strong> main experiment). This<br />

made it necessary to obtain at least four position measurements per day at<br />

each camp.<br />

On somewhat less evidence a space scale <strong>of</strong> 100 km was selected for study.<br />

Assuming approximately linear changes in velocity over that distance, three<br />

measuring points suffice to estimate deformation. Additional points provide<br />

84


some redundancy. Four points were used in the central <strong>AIDJEX</strong> array: three<br />

camps (Caribou, camp 1; Blue Fox, camp 2; Snow Bird, camp 3) initially forming<br />

a triangle 100 km on a side with a fourth camp (Big Bear, camp 0) in the<br />

center.<br />

Typically, sea ice moves about 2 km in 12 hours (Fig. 1). Thus, measurements<br />

<strong>of</strong> geographical position accurate to 100 m or better will resolve the<br />

motion quite well. However, the relative position <strong>of</strong> one camp with respect to<br />

another typically changes by about one quarter <strong>of</strong> one percent <strong>of</strong> the distance<br />

between them. For camps 100 km apart, then, typical changes in relative positions<br />

are 250 m in 12 hours (Fig. 2). The design criterion for the positioning<br />

system was set at 510 m for relative positions.<br />

Estimates <strong>of</strong> ice motion were required to support other parts <strong>of</strong> the <strong>AIDJEX</strong><br />

field program, but the requirements did not make the sampling and accuracy<br />

criteria more severe. Briefly stated, these other requirements were: to<br />

provide geographical reference for the meteorological and oceanographic observations;<br />

to correct ocean current measurements by removing the velocity <strong>of</strong> the<br />

ice itself; to correct measurements <strong>of</strong> ice tilt by accounting for the acceleration<br />

<strong>of</strong> the ice; and to support aircraft and submarine operations.<br />

SYSTEM DESIGN<br />

In defining the system configuration we sought to combine high reliability<br />

with low maintenance and repair, improve accuracy in distance measurement,<br />

have automatic operation with standardized selection and acquisition <strong>of</strong> the<br />

maximum number <strong>of</strong> passes, obtain measurements <strong>of</strong> floe rotation, and keep costs<br />

as low as possible. Figure 3 is a block diagram <strong>of</strong> the NavSat equipment<br />

installed at each <strong>of</strong> the four camps. Hardware details are given by Wasilew<br />

and Vivian [1976].<br />

Simple, redundant, commercially proven hardware was used to achieve our<br />

objectives <strong>of</strong> reliability and low cost. Spare components for major systems<br />

were kept at the main camp so that a system could be repaired quickly by<br />

replacing an entire component. We sacrificed some geographic accuracy by<br />

using single-channel (400 MHz) receivers which cannot correct for refraction<br />

effects. On the other hand, because <strong>of</strong> the low cost <strong>of</strong> single-channel<br />

85


eceivers, we were able to have two receivers at each camp and thus gained<br />

reliability through redundancy. As a result, out <strong>of</strong> a total <strong>of</strong> 1400 operating<br />

days, equipment problems prevented collecting any data on only 7 days<br />

at one camp.<br />

With two antennas at each camp separated by about 100 m, a comparison<br />

<strong>of</strong> fixes taken with each antenna permitted continuous monitoring <strong>of</strong> system<br />

accuracy and floe rotation. Accuracy in measuring distance was enhanced<br />

by adding a local clock to the receivers to remove timing errors and by<br />

selecting and acquiring passes under computer control.<br />

TRANSLOCATION<br />

For a stationary or slowly moving receiver, the largest error sources<br />

for single-channel positioning are uncorrected refraction and imperfect predic-<br />

tion <strong>of</strong> the satellite orbit.<br />

These errors are correlated between receivers<br />

tracking the same satellite for the same time interval.<br />

Therefore, the<br />

difference in position errors between the two receivers is much smaller than<br />

the errors in the position <strong>of</strong> each.<br />

h important condition is that the time intervals over which Doppler<br />

measurements are made at each station must be identical.<br />

we attempted to enforce this condition in real time.<br />

To avoid recomputing,<br />

interval for each pass to begin on the third even-minute mark before closest<br />

approach and to end on the third even-minute mark after closest approach,<br />

ieee9 10 minutes <strong>of</strong> data around the time <strong>of</strong> closest approach. When the times<br />

<strong>of</strong> closest approach for the two receivers fell in the same 2-minute interval,<br />

the critical intervals for the receivers were identical.<br />

Real-time<br />

We defined the cktieaZ<br />

translocation fixes were computed by using all data within the<br />

critical interval and eliminating all data outside it. Data collected during<br />

the IO-minute critical interval (20 Doppler counts, integrated for 30 seconds<br />

each) contain the essential features <strong>of</strong> the Doppler curve, so that discarding<br />

data from the beginning and end <strong>of</strong> the pass does little to degrade fix<br />

accuracy.<br />

(The critical interval concept was also used in the receiver con-<br />

trol algorithm; see the next section.)<br />

86


75.5<br />

75.:<br />

-<br />

E75.1<br />

LLI % y75.c -<br />

1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 ~<br />

* i -<br />

75.3 --,<br />

LONG I TU D E<br />

\%*<br />

f"<br />

75.2 - ! -<br />

1<br />

\<br />

- \<br />

-#+-,<br />

\ i #f- 'dd -<br />

LAT I TU DE<br />

r\ b-/ -<br />

k74.9 t, -<br />

74 -0<br />

-<br />

143.0<br />

144.0<br />

145 -0 E<br />

146.0<br />

F5LIJ<br />

0<br />

w<br />

0<br />

2<br />

c<br />

z 0 A<br />

74.7 --f-<br />

II km<br />

7-<br />

14 km<br />

147.0<br />

148-0<br />

-<br />

+ $. b a\<br />

r.2<br />

210-<br />

CI<br />

E 5'yV'::!Jf i;;.,:\,. .<br />

0.<br />

s 8s .:<br />

a<br />

Fig. 2. Distance between 205-<br />

'f *<br />

f;<br />

9.<br />

stations 2 (Blue Fox) t<br />

0 - -:; \<br />

and 3 (Snow Bird) from<br />

c,200r$<br />

unsmoothed NavSat<br />

U<br />

5<br />

translocation fixes. a.<br />

Pronounced diurnal % 195-<br />

oscillations are due<br />

I<br />

to inertial periods<br />

in the ice motion. 190<br />

'3;<br />

- :'5 yp:<br />

'1<br />

1 . 1 . 1 . 1 . 1 . 1 . 1 . 1 . 1 . 1 .<br />

N<br />

87


~~<br />

Ant cmiiii "A"<br />

(CHU- 319 2 )<br />

c<br />

1 1<br />

400 tlilz Receiver "A"<br />

(SPC scs-100)<br />

1- 5 Nlz Quart/<br />

O\ciLlator<br />

(Ausi ron 1250)<br />

7---<br />

__t___l<br />

]a1 Clock 1<br />

--Ip-+<br />

Antc,nna "4"<br />

(CHU- 319 2 )<br />

400 Mltz Rrceiver "D"<br />

(SPC scs-100)<br />

I<br />

1.[-I6<br />

L<br />

I<br />

1<br />

K Digital Computer<br />

(Data General 2/10) 1 Environmental I<br />

I<br />

Transport<br />

(Mangico Mod 7)<br />

I<br />

Fig. 3. NavSat equipment at each <strong>AIDJEX</strong> camp. Two singlechannel<br />

receivers share a quartz crystal frequency standard,<br />

antennas separated by about 100 m. The receiver acquisition<br />

circuits are controlled by a computer that also computes<br />

NavSat positions and logs environmental data as well as raw<br />

and processed NavSat data on magnetic tape.<br />

Fig. 4. A good receiver control procedure<br />

avoids a pass (A) that interferes<br />

with another pass (B) during<br />

3<br />

their critical intervals (solid 2<br />

lines); directs the search frequency<br />

$<br />

to a calculated value to acquire a 0<br />

pass (c) at a certain time; and w<br />

a<br />

break the receiver from a pass (D)<br />

and the end <strong>of</strong> its critical interval<br />

so that another pass (E) can be<br />

~r,<br />

0<br />

I/<br />

I I +


The translocation principle removes most <strong>of</strong> the errors external to the<br />

receivers. Still remaining are instrumentation errors at each receiver due<br />

to oscillator and receiver timing errors. For most <strong>of</strong> the main experiment,<br />

the crystal oscillators used in our systems introduced no detectable errors<br />

in our position measurements. However, because <strong>of</strong> inadequate oscillators,<br />

translocation calibration tests before the main experiment produced few useful<br />

data. During the first month, temperature fluctuations <strong>of</strong> a few tenths <strong>of</strong> a<br />

degree Fahrenheit in 10 minutes at the oscillators produced frequency changes<br />

<strong>of</strong> 1 part in lo1 O .<br />

Such a frequency drift will cause fix errors <strong>of</strong> 5-10 m<br />

[Denzler, 19701. Oscillators less sensitive to temperature were used after<br />

the first month.<br />

NavSat receivers used for navigation generally employ the simplification<br />

<strong>of</strong> measuring Doppler counts with respect to time marks transmitted by the<br />

satellite. Errors in decoding these time marks introduce 5-10 m position<br />

errors. Receivers intended for more precise applications eliminate this<br />

source <strong>of</strong> error by adding a stable internal time base to control or correct<br />

Doppler counts. We used such a local clock, derived from the crystal oscillator,<br />

to measure the errors in decoding satellite time marks and to correct<br />

Doppler counts using the algorithm in Appendix A.<br />

RECEIVER CONTROL<br />

There were six NavSat satellites during the experiment, each completing<br />

one orbit every 108 minutes and staying within range <strong>of</strong> high-latitude receivers<br />

for about 18 minutes on each orbit (Fig. 4). Therefore, it commonly occurred<br />

that more than one satellite was within range at the same time. Because our<br />

objectives included the collection <strong>of</strong> data from as many passes as possible<br />

while tracking the same passes at different camps, the standard practice <strong>of</strong><br />

tracking the first satellite within range for as long as possible was not<br />

efficient enough for us. The data collection had to be more selective and<br />

had to employ identical techniques at all camps.<br />

The objective <strong>of</strong> our receiver control procedure was to obtain data from<br />

the entire critical interval <strong>of</strong> as many passes as possible. The procedure<br />

involved logical choices based on predictions <strong>of</strong> Doppler frequencies as a<br />

89


function <strong>of</strong> time, control <strong>of</strong> the receiver search frequency, and the ability<br />

to break <strong>of</strong>f from a satellite after the end <strong>of</strong> the critical interval.<br />

From a total <strong>of</strong> 81 passes per day, generally 30-50 were chosen to be<br />

tracked. Only 50%-70% <strong>of</strong> the passes selected were actually tracked, due to<br />

shortcomings in the control <strong>of</strong> the receivers. The principal difficulty was<br />

interference <strong>of</strong> unwanted signals while trying to position the receiver<br />

frequency at the frequency <strong>of</strong> the desired satellite. However, early breaking<br />

from passes probably increased the number <strong>of</strong> fixes over what would have been<br />

expected from conventional automatic acquisition receivers. Proper implementation<br />

<strong>of</strong> the concept should result in nearly 100% success in tracking the<br />

desired passes and would increase the quantity and quality <strong>of</strong> data collected<br />

in high latitudes, especially during clustering <strong>of</strong> satellite passes.<br />

The average number <strong>of</strong> fixes from successfully tracked passes was 30 per<br />

day (Fig. 5). For each pair <strong>of</strong> stations, we typically acquired 20 translocation<br />

fixes per day.<br />

c<br />

Q,<br />

Fig. 5. Typical distribution <strong>of</strong><br />

fixes per day from a receiver<br />

for the full year <strong>of</strong> the main<br />

experiment. The histogram 0<br />

includes effects <strong>of</strong> system<br />

down-time, imperfect<br />

><br />

receiver control, and vari- 0<br />

ations in clustering <strong>of</strong><br />

satellite passes.<br />

3<br />

0<br />

W<br />

& 20<br />

5 IO<br />

a<br />

LL<br />

'0 10 20 30 40 50 60 70<br />

GOOD FIXES PER DAY<br />

90


POSITION ACCURACY<br />

To evaluate the quality <strong>of</strong> the position measurements, a period was chosen<br />

for analysis in which little or no ice motion occurred. During that period,<br />

14-24 February 1976, 500 fixes were taken at camp 1 (Caribou) and 500 at<br />

camp 3 (Snow Bird). From the position <strong>of</strong> camp 3 and the distance between<br />

camps as a function <strong>of</strong> time (Figs. 6 and 7), we deduce that ice motion was<br />

less than 20 m for the period.<br />

For each fix the radial error is defined as the vector from a reference<br />

position to the fix position. Stand-alone results refer to fixes from a<br />

single receiver (i.e., geographic accuracy). The reference position for these<br />

is taken to be the ensemble median position defined by the ensemble median<br />

W<br />

W<br />

g73-699<br />

,* LONGITUIIE .<br />

W<br />

0<br />

'73 =698<br />

W<br />

0<br />

E 73 696 4<br />

I- I50 m<br />

73.695<br />

..<br />

.. . ...<br />

. *<br />

i<br />

-144 =720<br />

-144.725 -<br />

-!44.730 5<br />

W<br />

0<br />

-144 a735<br />

0<br />

3<br />

-144.740 L<br />

(3<br />

z<br />

-144 -750<br />

t<br />

-144 -755<br />

LATITUDE<br />

73-692<br />

I I I I 1 I I I I I I 1 I ' '-144 -760<br />

73 -690'<br />

15 I7 19 21 23 25 27<br />

FEBRUARY 1976<br />

Fig. 6. Raw position fixes and smoothed filter output for station 3 during<br />

period used to evaluate position accuracy.<br />

91


102 .Q<br />

CRRIBOU TO SNOW BIRD<br />

-<br />

E<br />

Y<br />

Y<br />

W<br />

g 100.0<br />

!<br />

..<br />

*-!e--?-<br />

90 a 0<br />

6 8 10 12 14 16 18 20 22 24<br />

FEBRUARY I976<br />

Fig. 7. Distance between station 1 (Caribou) and station 3 (Snow Bird)<br />

from raw translocation fixes during period used to evaluate positioning<br />

accuracy.<br />

<strong>of</strong> the latitudes and longitudes.<br />

A translocation fix is defined as the vector<br />

difference <strong>of</strong> two fixes, one for each <strong>of</strong> two receivers. The translocation<br />

reference is the vector difference between the reference positions for each<br />

receiver antenna.<br />

A translocation radial error is the vector from the trans-<br />

location reference to the translocation fix.<br />

Our convention is to measure<br />

position accuracy as the 68th percent point <strong>of</strong> the lengths <strong>of</strong> the radial<br />

error vectors.<br />

Thus, an accuracy <strong>of</strong> 100 m for an ensemble <strong>of</strong> 500 fixes implies<br />

that exactly 340 fixes fall within a circle centered at the ensemble median<br />

position and or radius 100 m.<br />

Stand-A1 one Accuracy<br />

The stand-alone accuracy for the two ensembles <strong>of</strong> fixes is given in<br />

Table 1 and shown for camp 3 in Figure 8.3. These data are characterized as<br />

single-channel (no refraction correction), real-time, time-recovery-corrected<br />

data edited to conform with<br />

the critical interval described earlier.<br />

92


TABLE 1<br />

STAND-ALONE ACCURACY<br />

68th Percent Point<br />

Latitude Lonzitude Radial<br />

Camp 1 (Caribou) 30 m 55 m 67 m<br />

Camp 3 (Snow Bird) 28 m 62 m 69 m<br />

- Trans1 ocati on Accuracy<br />

The fixes from two receivers were scanned to find pairs <strong>of</strong> fixes from<br />

the same satellite with closest approach times in the same two-minute interval.<br />

This guarantees that the truncated fixes use data from exactly the same time<br />

interval.<br />

Table 2 and Figures 8b and 8c show the accuracy <strong>of</strong> translocation<br />

between camp 3 (Snow Bird) and camp 1 (Caribou) and between receivers A and<br />

B at camp 1 (Caribou).<br />

TABLE 2<br />

TRANSLOCATION ACCURACY<br />

Number<br />

Radial Error<br />

Receivers <strong>of</strong> Fixes Separation (68th % Point)<br />

1 - 3 199 98 km 20 m<br />

1A - 1B 241 94 m 5 m<br />

The translocation effect reduces the radial errors <strong>of</strong> about 70 m stand-<br />

alone at each camp to translocation errors <strong>of</strong> 20 m over about 100 km, and 5m<br />

over about 100 m.<br />

The improvement occurs because refraction errors and orbit<br />

prediction errors in the two fixes are highly correlated.<br />

coefficient p can be estimated as<br />

translocation error = 0.86 for 98 km<br />

= 2(stand-alone error) 0.96 for 94 m.<br />

The correlation<br />

The increase in radial error from 5 m to 20 m is probably due to ice motion<br />

and uncorrelated refraction at the 98 km separation.<br />

93


I<br />

NO8TH<br />

+<br />

A<br />

+ +<br />

8a (left). Stand-alone errors are<br />

dominated by uncorrected refraction<br />

effects, with a smaller<br />

contribution due to orbit prediction<br />

errors.<br />

STAND ALONE<br />

Q 367m<br />

t<br />

+<br />

I<br />

8b (right). The 100 km translocation<br />

errors are smaller because<br />

the orbit prediction. errors and<br />

most <strong>of</strong> the refraction effects are<br />

the same for the two receivers.<br />

I<br />

I<br />

I<br />

LONG RANGE<br />

TRAESLOCATIQN<br />

a=20m<br />

I<br />

t<br />

NORTH<br />

I<br />

50m ,<br />

I<br />

I<br />

+ + '4<br />

SHORT RANGE<br />

TRANS LO C A T I 0 tJ<br />

+ + *<br />

I + +<br />

I<br />

!<br />

EAST<br />

8c (left). The contribution <strong>of</strong><br />

instrumentation errors to translocation<br />

errors is seen in the<br />

100 m translocation.<br />

Fig. 8. Stand-alone and translocation positions from NavSat measurements from<br />

<strong>AIDJEX</strong> camps, 14-24 February 1976, a period <strong>of</strong> little ice motion. Circles<br />

contain 68% <strong>of</strong> the fix errors.<br />

94


Az i mu t h s<br />

When both receivers at a camp recorded a trunmttd fix from the same<br />

satellite, the azimuth and distance separating the two antennas were computed.<br />

There were about a dozen such NavSat azimuth determinations each day for each<br />

camp and, weather permitting, one celestial azimuth measurement. The celestial<br />

azimuths were taken with a theodolite and were accurate to 20.1'. The<br />

NavSat azimuths had 68th percentiles <strong>of</strong> about +2', and biases from 0' to 3'.<br />

Figure 9 is an example <strong>of</strong> NavSat and celestial measurements <strong>of</strong> floe rotation.<br />

250 '<br />

+ .?k<br />

+,<br />

+:-*<br />

.-r ........ - ..-.<br />

+ .- -*: ,:: . ...........<br />

Y<br />

'9<br />

u<br />

r... P.. - .".<br />

*<br />

9<br />

++. . -.<br />

*.-<br />

.+: ><br />

cc<br />

+ ......... .<br />

. .<br />

0<br />

'*<br />

-a<br />

w<br />

. " +. .... . -.. m<br />

0<br />

* - c.<br />

... + .. +<br />

..&*++ .. ..7<br />

' + ................ 2<br />

.t +*. .-. .. - ++ .... - - . .".. - -E<br />

...........<br />

I-<br />

'-1 ..<br />

2- ...... * a<br />

..... *<br />

-I<br />

. ?+:. -! .) -a c(<br />

. .-<br />

b<br />

v3<br />

....<br />

......<br />

.<br />

Y<br />

-8<br />

... ...<br />

.. -<br />

.<br />

Fig. 9. Floe rotation as measured by NavSat and celestial methods for camp 3<br />

(Snow Bird) during July 1975. Significant bias 1s- present in the NavSat<br />

measurements.<br />

95


VELOCITY ERROES<br />

As mentioned earlier, the February data were selected for detailed study<br />

because the ice was nearly stationary then. Clearly, such data do not include<br />

fix errors due to ice velocity itself. In the general situation when<br />

the ice is moving there is an additional source <strong>of</strong> error.<br />

The measurement equation which we used, relating the observer’s position<br />

to the corrected Doppler counts, did not take into account velocity <strong>of</strong> the<br />

observer. In principle the velocity can easily be included in the measurement<br />

equation--in fact, it is standard practice for shipboard applications--<br />

but we decided to ignore the velocity effect for several reasons. Ice<br />

velocities are typically small and highly correlated between stations only<br />

100 km apart. Thus the velocity errors in translocation will be small.<br />

Also, we were not confident <strong>of</strong> a technique to estimate the ice velocity based<br />

on the previous few fixes. The effect <strong>of</strong> velocity errors could have been<br />

reduced by a suitable algorithm to, estimate the ice velocity and use it in<br />

the fix solution.<br />

The histogram <strong>of</strong> ice velocity (Fig. 10) shows that the velocity error<br />

could occasionally exceed 20 cm sec-’ , with which, using the above guideline,<br />

is associated a fix error <strong>of</strong> at least 100 m. The difference in velocity<br />

between two stations will affect our translocation results. The velocity<br />

difference was occasionally as large as 10 cm sec-’ (FiF. I), causing occasional<br />

translocation errors <strong>of</strong> more than 50 m. Thus, we can see that at times<br />

<strong>of</strong> extreme ice motion, the effect <strong>of</strong> velocity errors is somewhat larger than<br />

the other sources <strong>of</strong> error in stand-alone and translocation position<br />

measurements.<br />

The errors in position due to velocity errors are roughly proportional<br />

to the ice velocity itself (50 m per 10 cm sec-l). Likewise, the distance<br />

traveled by the ice in a given tine (12 hours, say) is also proportional to<br />

the velocity. Thus, the ratio <strong>of</strong> the measurement error to the actual motion<br />

is independent <strong>of</strong> the ice velocity, and is about 0.02. While position errors<br />

due to the ice velocity are not negligible, they degrade our estimates <strong>of</strong> ice<br />

motion over periods <strong>of</strong> 12 hours or so by only approximately 2%.<br />

96


DATA PROCESSING<br />

Following the field experiment, the data from each camp were assembled<br />

for further processing. Only fixes satisfying the truncated fix criterion<br />

were used. These fixes were scanned to weed out fixes which were clearly in<br />

error by comparison with the rest <strong>of</strong> the data. The algorithm used for this<br />

purpose compared each fix with the median position <strong>of</strong> the 10 fixes preceding<br />

and the 10 fixes following it. On the first pass through the data, a fix was<br />

rejected if it differed from the median value by more than 10 km.<br />

second pass, a fix was rejected if it differed by more than 1 km.<br />

On the<br />

Azimuth<br />

fixes were treated in the same way, an azimuth measurement being rejected if<br />

it differed by more than 10' from the median value.<br />

The raw position measurements were processed using smoothing techniques<br />

given by Kalman and Bucy [1961] to produce estimates <strong>of</strong> the position, velocity,<br />

and acceleration <strong>of</strong> each station. The problem was formulated in terms <strong>of</strong> a<br />

state vector Xn at time t,, and a measurement vector 2,. X, contains the<br />

CARIROU-SNOWBIRD<br />

APRIL I975 - idAY 1976<br />

SEPARATION - I20 km<br />

f 4- W<br />

a c 20<br />

3 0 ~<br />

Fig. 10. Velocity and<br />

velocity differences<br />

occurring during the<br />

main experiment.<br />

t<br />

K<br />

Q,<br />

-<br />

L l<br />

OO 5 IO 15 20<br />

VELOCITY DIFFEHENCE (cm sect)<br />

*On<br />

CAR I BOU<br />

APRIL 1975-MAY 1976<br />

I<br />

-L<br />

OO 5 10 15 20<br />

VELOCITY (cm sec-'1<br />

0)<br />

IO<br />

a<br />

97


unknown position, velocity, and acceleration <strong>of</strong> each station.<br />

system as described by X is assumed to satisfy<br />

The physical<br />

.Yn+l = 4 (n+l ,n> xn + r (n+l ,n) w, ,<br />

where<br />

Here 4 and r are known matrices involving (t,+l - t,) and W, is an unknown<br />

random white noise. The covariance <strong>of</strong> w is assumed to be &.<br />

Z, contains the measured coordinates <strong>of</strong> each station which obtained a<br />

fix from the nth satellite pass. The measurements satisfy<br />

where<br />

Z, = Hn Xn -+ V,,<br />

‘in is the known relationship between the 2, and the Xn, and Vn is unknown<br />

random white measurement error with known covariance 8. Under these assump-<br />

tions, one can find best estimates €or X,<br />

given a sequence <strong>of</strong> measurerents<br />

52, i = 1,2, ..., €17 (see Thorndike and Cheung [1377] for details ).<br />

Of particular interest here is the definition <strong>of</strong> the measurement error<br />

covariance matrix R.<br />

Consider the situation in which two stations A and B<br />

tracked the same satellite pass. Then the measurement Z (looking at only one<br />

coordinate direction) has the form<br />

For times that the ice was moving, the estimation errors are up to twice as<br />

large as shown, since the raw data then include additional errors due to<br />

neglecting the ice velocity.<br />

Other data processing approaches could produce better accuracy with<br />

additional effort: treatment <strong>of</strong> velocity effects; use <strong>of</strong> passes tracked by<br />

the second receiver at each camp in position estimates; editing <strong>of</strong> the translocation<br />

data as well as the geographic data; incorporating models for the<br />

various error functions; and tighter tuning <strong>of</strong> the physical parameters that<br />

98


model ice motion in the statistical filter, We do not believe that our<br />

methods <strong>of</strong> data reduction provide the greatest accuracy? but they are<br />

expedient and they meet our needs.<br />

ACKNOWLEDGMENTS<br />

We gratefully acknowledge our introduction to the Navy Navigation<br />

Satellite System by the NavSat community, and the efforts <strong>of</strong> the employees<br />

<strong>of</strong> the Satellite Positioning Corporation, who built the positioning systems.<br />

Logistic support was provided by the Naval Arctic Research Laboratory and<br />

by the Distant Early Warning System in Alaska. We are indebted to the men<br />

and women who operated and maintained the equipment for one year on the<br />

Arctic Ocean. The work was funded by the National <strong>Science</strong> Foundation? the<br />

Office <strong>of</strong> Naval Research, and the Bedford Institute <strong>of</strong> Oceanography.<br />

APPENDIX A<br />

DOPPLER COllNTING CORRECTIONS<br />

The desired Doppler count is the integral <strong>of</strong> the Doppler frequency over<br />

the interval (tl, t2).<br />

<strong>of</strong> timing marks transmitted by the satellite at known times (T, T+T).<br />

tice the Doppler count is not measured exactly.<br />

The counting logic requires<br />

some time to recognize the timing marks during which time several full Doppler<br />

cycle marks may have passed.<br />

1.2 msec, with jitter <strong>of</strong> the order 50 usec.<br />

interval over which the Doppler measurement was made will generally be lagged<br />

with respect to the desired interval and <strong>of</strong> different length.<br />

An additional<br />

random error <strong>of</strong> 0-1 Doppler count at the beginning and end <strong>of</strong> each interval<br />

exists because the counting logic recognizes only zero crossings <strong>of</strong> the<br />

Doppler signal.<br />

follows.<br />

This interval is defined by the arrival at the antenna<br />

In our hardware? delays were estimated to be<br />

In prac-<br />

Figure 11 illustrates that the<br />

A corrected Doppler count B was constructed from the measured value D as<br />

The Doppler count can be interpreted as an area under the curve <strong>of</strong><br />

Doppler frequency versus time (Fig. 11).<br />

Therefore,<br />

99


l+b 1 ,t 2+b 2<br />

The right-hand side can be rewritten as<br />

-<br />

where b = (bl + b2)/2 is the average delay.<br />

difference between the measured and the desired interval.<br />

The desired<br />

interval length can be estimated using the fundamental Doppler navigation<br />

equation:<br />

A<br />

D = (Po - F s ) ~<br />

change in slant range<br />

+ Fo<br />

speed <strong>of</strong> light ’<br />

The quantity (bl - b2) is the<br />

where Fs and F, are the frequencies <strong>of</strong> the satellite and receiver oscilla-<br />

tors (roughly 400 MHz), and T is 120x1526/6103 seconds for the first three<br />

counting intervals in each 2-minute period and 120x1525/6103 seconds for<br />

the fourth.<br />

The duration <strong>of</strong> the measured interval was determined by counting<br />

the cycles <strong>of</strong> the receiver clock during the interval between the first<br />

DOPP<br />

whe r<br />

t2+b,.<br />

er zero crossing <strong>of</strong> successive counting intervals:<br />

N<br />

measured interval length = -<br />

Fo<br />

N is the number <strong>of</strong> cycles at the navigator frequen c F, from t,+b, to<br />

The equation for the corrected Doppler count is<br />

where the 5 term corrects for average delay and the last term corrects for<br />

jitter. In this expression fi is the Doppler frequency at ti estimated by<br />

counting Doppler cycles di during the first 20 msec after t;+bi:<br />

fi = di/2Q msec.<br />

The typical magnitudes <strong>of</strong> the corrections are 0.3-1.8 counts for delay<br />

and about 1.5 counts for length. This correction reduces the error in the<br />

corrected Doppler counts to less than 0.1 count.<br />

100


FREC<br />

1<br />

INCY<br />

/<br />

DOPPLER<br />

/<br />

FREQUENCY<br />

/<br />

----4<br />

MEASURED<br />

L TIME<br />

/<br />

DESIRED INTERVAL<br />

Fig. 11. The Doppler count is an intestral <strong>of</strong> the<br />

Doppler frequency versus time, an area under the<br />

curve plotted here. The timing delays b, and b,<br />

shift the measurement interval and thus alter the<br />

area. A correction--the "time recovery" correction--is<br />

described in the text.<br />

RE FEREN CES<br />

Brown, W. P., and E. G. Kerut, 1975. Arctic environment buoy system. In<br />

Ocean 75, Proceedings 19 75 IEEE InternationaZ Conference on Engineering<br />

in the Ocean Environment, IEEE Publ. No. CHO 995-1 OEC, pp. 50-55.<br />

Also in this <strong>Bulletin</strong>, pp. 15-20.<br />

Denzler, D. W. R. 1970. Translocation Reference Manual. Rep. No. S3-R-007<br />

(Oct. 1970), Johns Hopkins <strong>University</strong>, Applied Physics Laboratory, Silver<br />

Springs, Md.<br />

Kalman, R. E., and R. S. Bucy. 1961. New results in linear filtering and<br />

prediction theory. Journal <strong>of</strong> Basic Engineering, 830, 95-108.<br />

Thorndike, A. S. 1974. Strain calculations using <strong>AIDJEX</strong> 1972 position data.<br />

<strong>AIDJEX</strong> <strong>Bulletin</strong>, 24, 107-129.<br />

Thorndike, A. S., and J. Cheung. 1977. <strong>AIDJEX</strong> measurements <strong>of</strong> sea ice<br />

motion, 11 April 1975 to 14 May 1976. <strong>AIDJEX</strong> <strong>Bulletin</strong>, 35, 1-149.<br />

Wasilew, D., and J. Vivian. 1976. Satellite positioning techniques in the<br />

Arctic. In Proceedings Eighth Annual Offshore Techno logy Conference,<br />

paper no. OTC 2460, Dallas, Texas.<br />

101


AIRCRAFT LOGISTICS AND DRIFTING BUOY COVERAGE<br />

FOR THE FIRST GARP GLOBAL EXPERIMENT<br />

Pat Martin<br />

<strong>AIDJEX</strong><br />

There has been a shift <strong>of</strong> emphasis in the plans for distribution <strong>of</strong><br />

drifting data buoys in FGGE away from the Antarctic convergence and toward<br />

the tropics in response to the elimination <strong>of</strong> the carrier balloon/dropsonde<br />

program from the observing system. Early plans had called for 300 buoys to<br />

be deployed between latitudes 55's and 65's to collect surf ace temperature<br />

and pressure data not available from satellite soundings due to the persistent<br />

cloudinessin that region. Recent deployment plans emphasize coverage<br />

<strong>of</strong> temperate latitudes (3OoS-55'S) and have very sparse coverage south <strong>of</strong><br />

55OS (Fig. 1). The hope <strong>of</strong> polar scientists to have significant new data in<br />

conjunction with FGGE is not reflected in FGGE data buoy plans.<br />

The schedule for buoy deployments reflects a heavy emphasis on the first<br />

special observing period (SOP) in the austral summer (Fig. 2). Of the 240<br />

buoys covered in the current plans, 185 are scheduled to be deployed before<br />

May 1979 and 55 after that. The few deployments in May and June which could<br />

be effective in producing good coverage <strong>of</strong> the effects <strong>of</strong> the Antarctic winter<br />

on the Southern Ocean are, instead, to be deployed primarily between 20's and<br />

4OoS, as seen in the hatched area <strong>of</strong> Figure 1. In this respect the FGGE is<br />

rapidly becoming a rare opportunity lost to polar science. For the first<br />

(and perhaps only) time the observing system in the Southern Hemisphere will<br />

be reasonably complete, except for the ocean south <strong>of</strong> the Antarctic convergence<br />

and especially in the austral winter.<br />

This unfortunate situation came about in part because polar scientists<br />

were not aware <strong>of</strong> the extent to which the shift to lower latitudes would<br />

dilute high latitude coverage and did not press decisions for a higher priority<br />

on polar coverage or establish independent programs. The shift <strong>of</strong><br />

emphasis to mid-lat5tudes and summer was also necessitated by the availability<br />

103


35<br />

30<br />

25<br />

v)<br />

><br />

0<br />

2 20<br />

LL<br />

0<br />

r=<br />

E m 15<br />

I<br />

3<br />

z<br />

IO<br />

JANUARY SOP<br />

/<br />

1<br />

A DEPLOYMENTS<br />

I 60-<br />

I-<br />

z<br />

0<br />

x<br />

I 50-<br />

0<br />

a<br />

W<br />

n<br />

W 40-<br />

><br />

0<br />

J<br />

a<br />

W<br />

n 30-<br />

v)<br />

><br />

0<br />

3<br />

20-<br />

iL<br />

0<br />

5<br />

0<br />

Figure 1.<br />

25<br />

35 45 55 65<br />

LATITUDE<br />

Estitiiated distribution <strong>of</strong> buoys with latitude south for the first<br />

and second special observing periods.<br />

tude south for the second special observinq period are also shown in the hatched<br />

area.<br />

Planned deployments <strong>of</strong> buoys with lati-<br />

The plan elliphasizes mid-lati tudes and sur.ir:ier deploymnts.<br />

W<br />

m<br />

L<br />

3<br />

z<br />

2 3 4 5 6<br />

1978 1 2 1 ' 1979<br />

MONTHS<br />

Figure 2. The schedule for buoy deployments showinq heavy<br />

emphasis on summer deployment.


<strong>of</strong> ships-<strong>of</strong>-opportunity, the primary method <strong>of</strong> deployment. There are few<br />

enough ships south <strong>of</strong> the Antarctic convergence in the summer, and virtually<br />

none in the winter. The emphasis on contributions <strong>of</strong> ship time from participating<br />

countries to deploy buoys, rather than pushing for equivalent<br />

contributions <strong>of</strong> aircraft time, reflects the strong ship-based oceanographic<br />

background <strong>of</strong> most persons involved in data buoy developments. This has<br />

placed unnecessary restrictions on the strategies for data buoy deployments<br />

in FGGE.<br />

Aircraft deployment is essentially the only choice available in the<br />

Arctic Ocean, and has been in use for more than 20 years. Parachute deployment<br />

<strong>of</strong> buoys has been an important recent development, with about 50 buoys<br />

deployed without a failure in the past two years. Several successful tests<br />

<strong>of</strong> parachute deployment <strong>of</strong> drifting buoys have been made into open water,<br />

including the successful operation <strong>of</strong> thermistor strings and barometric pressure<br />

sensors after such drops. Both open ocean spar buoyand buoys that sit<br />

on top <strong>of</strong> the ice have worked for as much as a full year. Controlled tests<br />

<strong>of</strong> the ice-sitting version have shown pressure measurements better than 0.3<br />

mb and daily average temperature measurements within l0C over a four-month<br />

period. Parachute deployment <strong>of</strong> buoys has been shown to be a practical technique.<br />

The Antarctic continent, and the area covered by sea ice that surrounds<br />

it, is the most difficult place on the globe to fly to in the winter. Most<br />

<strong>of</strong> the continent, and the entire ocean area surrounding it, can be reached<br />

from four major southern hemisphere cities using an aircraft with 4500 km<br />

(2700 n. mi.) operating radius (Fig. 3). With a standard fuel load, the<br />

C-141 Starlifter, a jet cargo aircraft, can operate at this radius with a<br />

650 km (400 n. mi.) reserve. The C-130 Hercules, a turboprop cargo plane in<br />

commercial and government service throughout the world, can operate to a<br />

radius <strong>of</strong> 3000 km (1800 n. mi.) with a 500 km (33 n. mi.) reserve. Operating<br />

from a smaller airstrip at the tip <strong>of</strong> South America, a C-130 could reach the<br />

same locations near Antarctica as a C-141 flying from Buenos Aires. Similarly,<br />

there are other areas throughout the world where smaller aircraft can be used,<br />

sometimes out <strong>of</strong> more remote airstrips, to deploy data buoys where and when<br />

they are needed.<br />

105


An example <strong>of</strong> a large-scale buoy array planned for deployment by aircraft<br />

during the FGGE is shown in Figure 4. This array <strong>of</strong> 20-30 buoys would<br />

be deployed in the fall <strong>of</strong> 1978 using four C-130 flights. The availability<br />

<strong>of</strong> airstrips in Canada and Greenland closer to the Arctic Basin than Thule<br />

makes the majority <strong>of</strong> the deployments feasible from smaller aircraft such as<br />

the DHC-6 Twin Otter.<br />

When a major purpose <strong>of</strong> a buoy deployment is to collect surface barometric<br />

pressure data for a full year, there is good reason for concern about<br />

calibration drift <strong>of</strong> the sensor. It is always economic to invest in a sensor<br />

in which confidence can be placed for the duration without resorting to field<br />

checks. These is some evidence that we are approaching a time when several<br />

different sensors will merit this confidence. For the FGGE, where many different<br />

types <strong>of</strong> sensors are to be deployed by many different groups, it may<br />

be desirable to check the calibrations prior to the second special observing<br />

period. Again, aircraft <strong>of</strong>fer a means <strong>of</strong> carrying out this task efficiently,<br />

since the same large distances covered during deployments are involved in<br />

overflying buoys for calibration.<br />

Since, to achieve maximum range, turbine engines used in long-range<br />

aircraft must be operated at altitudes <strong>of</strong> 20,000-40,000 feet, it is impractical<br />

to consider low-level flights to measure surface pressure directly.<br />

This problem is compounded by the real difficulty <strong>of</strong> measuring static pressure<br />

from a fast-moving aircraft. Instead, the calibration can be made by<br />

dropping pressure sensors to the surface with radio telemetry to the aircraft.<br />

This technique is common for temperature sounding, which will be carried out<br />

during the FGGE over large areas <strong>of</strong> the tropical oceans from C-141 aircraft.<br />

The pressure sensor added to such dropsondes need not be expensive, since it<br />

would be expected to operate for less than an hour and through only one pressure<br />

excursion. Provision for survival <strong>of</strong> the dropsondes after water or sea<br />

ice landing would be necessary to allow the temperature <strong>of</strong> the sensor to<br />

stabilize so that several minutes <strong>of</strong> good data could be collected.<br />

Parachute deployment <strong>of</strong> data buoys, together with the proven performance<br />

<strong>of</strong> satellite data collection and tracking, aremaking dramatic changes in the<br />

nature <strong>of</strong> studies <strong>of</strong> Arctic air-ice-ocean interaction. It seems likely that<br />

aircraft deployments will play an increasingly important role wherever surface<br />

106


Figure 3. Coverage <strong>of</strong> the Southern Ocean possible<br />

within 4500 km (2700 n. mi.) <strong>of</strong> major southern<br />

helllisphere cities. The dots and circles represent<br />

buoy locations and deployment sites, respectively,<br />

as shown in the draft deployment plan.<br />

Figure 4. An example <strong>of</strong> flight tracks needed to<br />

b<br />

deploy a large-scale array in the Arctic Ocean.


observations are needed in remote areas. Purchase, deployment, and maintenance<br />

<strong>of</strong> the 500-1000 buoys necessary to complete the worldwide surface<br />

observing network require no new developments and could be carried out for<br />

about $5 million annual expense. The FGGE is the first opportunity for these<br />

developments to be used on a global scale, but by no means the last.<br />

108


MATHEMATICAL CHARACTERISTICS OF A PLASTIC MODEL<br />

OF SEA ICE DYNAMICS<br />

Robert S. Pritchard and R. Reimer<br />

<strong>AIDJEX</strong><br />

ABSTRACT<br />

A plastic sea ice model developed by the <strong>AIDJEX</strong> modeling group is<br />

analyzed to determine the conditions under which real characteristic<br />

curves exist. For this analysis, inertia and material<br />

hardening are assumed negligible. We show that the characteristics<br />

at each point where the material is plastic can be real and<br />

distinct (hyperbolic equations), coincident (parabolic equations),<br />

or imaginary (elliptic equations). There may also be elastic<br />

regions. The characteristics enable one to see which parts <strong>of</strong><br />

the boundary affect which parts <strong>of</strong> the solution region, and<br />

thereby they show where discontinuities in the solution may be<br />

expected.<br />

The characteristic curves do not depend on advection, air stress,<br />

water drag, Coriolis force, sea surface tilt, or yield strength<br />

gradients except as these terms affect the stress state. At each<br />

point the direction taken by the characteristic curves is determined<br />

as a function <strong>of</strong> the stress state. The curves are<br />

symmetric about the principal stress axes, while the angle<br />

between them depends on the position <strong>of</strong> the stress on the yield<br />

curve. The governing set <strong>of</strong> partial differential equations are<br />

transformed into ordinary differential equations along the characteristic<br />

curves. These equations have been determined for an<br />

arbitrary isotrspic yield surface and a non-normal flow rule when<br />

advection is included.<br />

The analytical difficulties that arise when a normal flow rule<br />

and advection are simultaneously considered are discussed. It<br />

has been conjectured that in many cases the large-scale lead<br />

patterns in the ice cover are related to the characteristic<br />

curves. Since velocity discontinuities are anticipated across<br />

leads it is the velocity characteristics at which we look. Along<br />

these curves there is no stretching. This physical property<br />

neither provides a physical explanation for the correspondence<br />

between leads and characteristics nor refutes such a possibility.<br />

Further observations are required. The characteristic analysis<br />

improves our understanding <strong>of</strong> the effect <strong>of</strong> yield surface and<br />

flow rule on ice response. The characteristic equations will be<br />

109


useful for determining how various parameters may control. ice<br />

response.<br />

INTRODUCTION<br />

It is well known that the ice cover <strong>of</strong> the polar seas deforms appreciably.<br />

The forces that cause this motion come fromthewinds and oceancurrents.<br />

During the past several decades there have been many attempts to develop a<br />

mathematical model that would allow this motion to be described and simulated.<br />

These models have addressed the problem on different time and length<br />

scales. We are interested in motions defined on length scales with resolution<br />

on the order <strong>of</strong> tens <strong>of</strong> kilometers and time resolution on the order <strong>of</strong><br />

one day. To understand the interaction <strong>of</strong> sea ice with its environment on<br />

these scales was the primary purpose <strong>of</strong> the <strong>AIDJEX</strong> program (Maykut et al.,<br />

1972). In the present work we shall confine our attention to the matheuatical<br />

properties <strong>of</strong> a special form <strong>of</strong> the model.<br />

The elasticfplastic representation was chosen by the <strong>AIDJEX</strong> modeling<br />

group because it is consistent with the processes that occur in the deformation<br />

<strong>of</strong> the ice cover. In particular, the mechanism <strong>of</strong> ridging, described<br />

by Parmerter and Coon (1972), has the properties that ridge formation is<br />

independent <strong>of</strong> the rate at which deformation occurs and that limit heights<br />

are achieved. Both <strong>of</strong> these properties are compatible with the plastic representation<br />

<strong>of</strong> the ridging process. Since a plastic model is independent <strong>of</strong><br />

the rate at which deformations occur, this model was also felt to be a suitable<br />

candidate for describing deformations in the ice cover that in the<br />

central Beaufort Sea can be approximately one percent per day and in the<br />

nearshore regions <strong>of</strong> the Alaska North Slope can be on the order <strong>of</strong> fifteen<br />

percent per day. Although the original model was developed to simulate the<br />

response <strong>of</strong> the ice cover observed during the <strong>AIDJEX</strong> experiment, the purpose<br />

and use <strong>of</strong> this model was expanded to include response <strong>of</strong> the ice cover near<br />

shore.<br />

To allow approximate integration <strong>of</strong> the system <strong>of</strong> coupled nonlinear<br />

partial differential equations, a numerical scheme was developed by Pritchard<br />

and Colony (1976). The s<strong>of</strong>tware was designed so that conditions observed<br />

110


during the <strong>AIDJEX</strong> main experiment could be simulated (Colony, 1975).<br />

addition to simulations <strong>of</strong> observed conditions, numerous calculations have<br />

been performed using idealized driving forces (Pritchard and Schwaegler,<br />

1975; Pritchard and Colony, 1974; Schwaegler and Pritchard, 1977). Complete<br />

simulations during special time periods in the 1975-76 <strong>AIDJEX</strong> main experiment<br />

also allowed a test <strong>of</strong> the model under realistic driving forces.<br />

last cases the model response was compared with observed response <strong>of</strong> the ice<br />

cover. These results were reported by Coon et al. (1976, 1977) and by<br />

Pritchard et al. (1976, 1977).<br />

In<br />

In these<br />

While the aforementioned calculations have increased understanding <strong>of</strong><br />

how the model responds to a given set <strong>of</strong> conditions, they have also raised<br />

some questions.<br />

The lack <strong>of</strong> answers is <strong>of</strong>ten caused by the difficulty <strong>of</strong><br />

understanding how each physical property affects response, 'and this diffi-<br />

culty is enhanced by nonlinearities <strong>of</strong> the model.<br />

are requi.red, it is particularly difficult to determine accurately special<br />

local features <strong>of</strong> the response, such as velocity discontinuities. Because<br />

<strong>of</strong> these difficulties, the present work was undertaken.<br />

out that the jump conditions that must be satisfied when discontinuities <strong>of</strong><br />

various kinds are present were discussed by Nye (1975).<br />

Since numerical solutions<br />

It should be pointed<br />

That analysis, how-<br />

ever, did not address the question <strong>of</strong> when or where discontinuites might<br />

arise or how solutions might vary along discontinuities.<br />

It is known that<br />

characteristics play a dominant role in the nature <strong>of</strong> the solutions <strong>of</strong> cer-<br />

tain systems <strong>of</strong> partial differential equations (Courant an& Hilbert, 1962).<br />

For example, discontinuities may exist in solutions only along characteris-<br />

tic curves. Since some partial differential equations do not admit real<br />

characteristic curves, these systems eliminate the possibility <strong>of</strong> a discont<br />

inuity .<br />

Satellite images <strong>of</strong> the ice cover (both NOAA-4 and Landsat) have shown<br />

that deformations are <strong>of</strong>ten concentrated in narrow bands that could be well<br />

approximated as velocity discontinuities on time scales <strong>of</strong> a day. Between<br />

these discontinuities, deformations <strong>of</strong> lower magnitudes also occur. It is<br />

our feeling that these features can be represented by a plastic ice model.<br />

Therefore, it is important to understand how, when, and where discontinui-<br />

ties may arise. Since discontinuities may arise in modeled stress fields<br />

111


as well as velocity fields, we must understand also how eachoccurs. However,<br />

the satellite images definitely show velocity, not stress, discontinuities<br />

and so it is this field on which our primary interest is focused.<br />

In summary, the present work is expected to provide useful information<br />

on several questions. Our intent is to study the mathematical characteristics<br />

<strong>of</strong> the <strong>AIDJEX</strong> model (1) to understandhow,where, andwhy discontinuities<br />

may appear, (2) to aid in the interpretation <strong>of</strong> the numerical solutions,<br />

(3) to learn how the material parameters affect the solutions, (4) to learn<br />

whether or not there are response features that can be observed in the ice<br />

cover, and (5) to determine equations along characteristic curves to allow<br />

development <strong>of</strong> analytical solution techniques.<br />

BACKGROUND<br />

Our interest in the characteristic analysis was aroused by the work <strong>of</strong><br />

Marco and Thompson (1975, 1977), who studied the regular pattern <strong>of</strong> leads<br />

that are sometimes observed over large portions <strong>of</strong> the Beaufort and Chukchi<br />

Seas. Their several attempts to explain the lead patterns have led to the<br />

conclusion that brittle material fracture is the cause. We, however, do not<br />

share their belief that plasticity cannot describe the phenomenon. Their<br />

analysis <strong>of</strong> mathematical characteristics associated with a plastic model<br />

depends on a von Mises yield surface and normal flow rule that prohibits<br />

dilatation. Theseare notproperties <strong>of</strong> the <strong>AIDJEX</strong> plastic model. It is<br />

felt that plasticity may be able to explain the lead patterns, and the<br />

present work does show such a possibility. However, the interpretation <strong>of</strong><br />

lead patterns in terms <strong>of</strong> characteristic directions by Marco and Thompson<br />

does point toward an analysis that extends our knowledge <strong>of</strong> properties <strong>of</strong><br />

the <strong>AIDJEX</strong> model and, consequently, <strong>of</strong> sea ice dynamics. Therefore, we consider<br />

their work to be <strong>of</strong> particular importance.<br />

The method <strong>of</strong> characteristics applied to a system <strong>of</strong> first order differential<br />

equations consists <strong>of</strong> finding the characteristic directions, or<br />

simply the characteristics, and a differential equation for each characteristic.<br />

When this equation is differentiated in only the characteristic<br />

direction it is called a characteristic relation. The system <strong>of</strong> differential<br />

112


equations is classified as hyperbolic, parabolic, or elliptic, depending on<br />

the number <strong>of</strong> real characteristic directions that exist. In plasticity<br />

problems this will be determined by the state <strong>of</strong> stress and the type and<br />

shape <strong>of</strong> yield surface that is used. In statically indeterminate problems<br />

dealing with materials described by a normal flow rule theremaybearepeated<br />

root in the characteristic determinant. To the authors' knowledge, this<br />

problem has not been dealt with in the plasticity literature. It is recognized<br />

by Martin (1975), who shows that the problem may be circumvented by<br />

neglecting the advection term in momentum balance. This has the effect <strong>of</strong><br />

uncoupling the stress equations from the velocity equations so that two<br />

independent systems may be solved.<br />

In the field <strong>of</strong> plasticity <strong>of</strong> metals, Hill (1950) solves problems with<br />

no body forces or advection. He discu6ses the possibility <strong>of</strong> parabolic and<br />

elliptic systems and obtains characteristic relations in the hyperbolic case.<br />

The well-known theorem due to Hencky for construction <strong>of</strong> slipline networks<br />

results from this assumption. Martin (1975) solves the same type<strong>of</strong>problems,<br />

but uses the null-vector technique, which is more easily generalized to<br />

include body forces and advection.<br />

The important distinction between plane strain and plane stress in<br />

plasticity is discussed by Hodge (1950), and Prager and Hodge (1951), as<br />

well as Hill (1950). It must be remembered that plane stress and strain are<br />

special cases <strong>of</strong> three-dimensional models, and our sea ice model is not<br />

derived from a three-dimensional model; it is a two-dimensional model.<br />

In plane strain the von Mises and Tresca yield conditions coincide,<br />

while in plane stress they do not. In plane strain the characteristics <strong>of</strong><br />

the stress equations exist at every point; they form an orthogonal network<br />

and are identical with the lines <strong>of</strong> maximum shearing stress and maximum<br />

shearing strain. In plane stress, if the Tresca yield condition is used<br />

together with the stress-strain law <strong>of</strong> von Mises, the stress equations may<br />

be hyperbolic in part <strong>of</strong> the plastic domain and parabolic in the rest.<br />

Characteristics <strong>of</strong> the velocity equations differ from those <strong>of</strong> the stress<br />

equations in this case. If the von Mises condition is used in plane stress,<br />

the stress and velocity characteristics coincide, but the system may be<br />

hyperbolic, parabolic, or elliptic.<br />

113


Since small changes in the yield surface may change the system <strong>of</strong> equations<br />

from hyperbolic to elliptic, it is <strong>of</strong> interest to know what the<br />

corresponding change in the actual stress state will be. Hodge (1950) has<br />

shown that a small change in a von Mises yield surface produces small changes<br />

in the stress field in the problem <strong>of</strong> an expanding circular hole in a plate,<br />

while for a notched bar in tension the stress state does not vary smoothly as<br />

the system goes from hyperbolic to elliptic.<br />

Problems in the plasticity <strong>of</strong> soils are solved by using a Coulomb yield<br />

function. They differ from metals problems where the materials are typically<br />

assumed incompressible. De Jong (1959) and Haithornwaite (1963) have argued<br />

that if this is to be the case, then a non-normal flow rule must be used.<br />

At the same time, Saint-Venant's hypothesis that the principal directions <strong>of</strong><br />

the stress and strain tensors coincide is found to give unrealistic results.<br />

The alternative, formulated by Shield (1955) and Jenike (1961), is to allow<br />

the cohesion <strong>of</strong> the soil to be a function <strong>of</strong> density and to allow workhardening.<br />

Spencer (1964) obtains equations governing the velocity field for an<br />

incompressible material, but does not integrate along the characteristic<br />

directions. Likewise, Morrison and Richmond (1976) extend Spencer's result<br />

to include body forces and solve the problem <strong>of</strong> gravity flow through a<br />

restricting channel. This solution, as with all solutions in soils plasticity,<br />

is for a state <strong>of</strong> plane strain. When the Coulomb yield criterion is<br />

used the possibility <strong>of</strong> non-real characteristic roots does not occur, so the<br />

equations are always hyperbolic. The problem <strong>of</strong> repeated characteristic<br />

roots is not encountered, either.<br />

Recently, Sodhi (1977) has taken the plane strain solution <strong>of</strong> Morrison<br />

and Richmond (1976) and applied it to the flow <strong>of</strong> pack ice througharestricting<br />

channel. By correlating lead directions in the icepackwithcharacteristic<br />

directions, he predicts the conditions required for ice breakout to occur<br />

and, by measuring the relative orientation <strong>of</strong> leads through a point, deduces<br />

the cohesive strength <strong>of</strong> the ice cover. This model differs both in the<br />

yield surface and the flow rule from the <strong>AIDJEX</strong> plastic model and the results<br />

are not applicable to our work.<br />

114


As can be seen from the references, the mathematical theory <strong>of</strong> characteristics<br />

is well established for many systems <strong>of</strong> equations, and much work<br />

has been done to define the characteristic curves for plasticity models.<br />

Some <strong>of</strong> the results that we present have been published previously. However,<br />

the lack <strong>of</strong> common notation and approach has made it difficult to understand<br />

whether or not a given result is applicable to our model. Therefore, we have<br />

introduced the model to be studied and have made the simpler classical analysis<br />

in a format that is consistent without analysis <strong>of</strong> the complete model.<br />

This approach has allowed us to understand the limitations and lack there<strong>of</strong><br />

in each simple case.<br />

Since the mathematical characteristics <strong>of</strong> the system <strong>of</strong> equations may<br />

depend on the particular form <strong>of</strong> the equations, we must discuss the particular<br />

form <strong>of</strong> the model to be considered. We choose to work with an ideal<br />

plastic form <strong>of</strong> the model so that the time variable does not enter the system<br />

<strong>of</strong> equations. It has been well documented that acceleration <strong>of</strong> the ice cover<br />

on temporal scales <strong>of</strong> one day are small enough that inertia may be neglected.<br />

With this assumption in a rigid plastic material model we have time entering<br />

the model only in the form <strong>of</strong> a hardening law. To eliminate this dependence,<br />

we have neglected hardening and assumed that hardening occurs at a slower<br />

rate.<br />

We consider both the existence <strong>of</strong> real characteristic curves and the<br />

form <strong>of</strong> the governing equations along these curves when the system is hyperbolic.<br />

The characteristic directions are discussed in detail and are<br />

related to solution variables <strong>of</strong> importance. Several examples <strong>of</strong> this relation<br />

are shown to help understand how the characteristic directions might be<br />

useful in interpreting the performance <strong>of</strong> the model. The analysis is performed<br />

for the most general system, which includes advection <strong>of</strong> the.materia1,<br />

but a non-normal flow rule is required to complete the analysis. In this<br />

system, stress and velocity variables are coupled. The flow rule is not<br />

normal, but it is defined from a potential function as an assumption that<br />

makes principal directions <strong>of</strong> stress and stretching coincide. The properties<br />

<strong>of</strong> this system are also examined when advection is neglected and when<br />

a normal flow rule is assumed. The problems that arise when a normal flow<br />

rule is assumed are discussed, but this analysis is not completed.<br />

115


The transformation <strong>of</strong> the system <strong>of</strong> equations into characteristic coordinates<br />

is presented by first discussing the cases where advection may be neglected<br />

so that the system <strong>of</strong> equations separates into two independent portions, one<br />

for stress and the other for velocity. Characteristic equations are then<br />

found for the general case where advection and the general flow rule are<br />

considered.<br />

MODEL<br />

The mathematical model developed by the <strong>AIDJEX</strong> modeling group to<br />

describe the response <strong>of</strong> the sea ice cover to the driving forces is presented<br />

by Coon et al. (1974). The forces contributing to the change <strong>of</strong> momentum in<br />

the horizontal plane <strong>of</strong> the ice cover are the air stress, water stress, internal<br />

ice stress divergence, Coriolis force, and sea surface tilt. The time<br />

and space scales <strong>of</strong> interest are approximately one day and tens <strong>of</strong> kilometers.<br />

On these scales it may be assumed that the acceleration occurring in the ice<br />

is small enough that inertia may be neglected in the momentum balance. We<br />

shall make this assumption in the current work so that the number <strong>of</strong> independent<br />

variables needed to define the equation may be limited to the two<br />

spatial coordinates. Were we not to consider a quasi-steady description <strong>of</strong><br />

the model, we would have to include time as an independent variable, thereby<br />

increasing the complexity <strong>of</strong> the characteristic analysis.<br />

The horizontal momentum balance may be written as<br />

where<br />

v<br />

-g<br />

:a<br />

2 is ice velocity,<br />

is geostrophic ocean current,<br />

is air stress,<br />

zw = - ) is water stress,<br />

g<br />

m is areal mass density,<br />

$ is the material rate <strong>of</strong> change <strong>of</strong> 2,<br />

V.0 is the ice stress divergence,<br />

116


fc is the Coriolis parameter,<br />

- k is a unit vector normal to the earth's surface;<br />

To study the quasi-steady problem, we should not neglect mg, but instead<br />

write:<br />

where & = ag/ag is the velocity gradient. The model is then written in an<br />

Eulerian description.<br />

Now it is meaningful to neglect av_/at so that time<br />

dependence may be ignored. With these assumptions in mind, we may write the<br />

Cartesian components <strong>of</strong> momentum balance as<br />

where Cartesian components <strong>of</strong> all tensors are introduced:<br />

and the water stress law may include any nonlinear algebraic response that<br />

depends only on the ice velocity relative to geostrophic ocean currents<br />

(McPhee, 1975).<br />

The constitutive law used to simulate sea ice dynamics is elasticplastic<br />

(Coon et al., 1974; Coon and Pritchard, 1974; Pritchard, 1975).<br />

However, the elastic response is included primarily to help obtain a numerical<br />

solution. A rigid-plastic material law is thought to be as reasonable<br />

from a physical standpoint. Our analysis <strong>of</strong> characteristics to date has<br />

treated only the rigid-plastic model. Therefore, we must exercise care when<br />

the results <strong>of</strong> this analysis are used to interpret possible discontinuities<br />

in numerically calculated results. But we shall study the mathematical<br />

properties <strong>of</strong> the rigid-plastic model, and from these results we expect to<br />

117


e able to judge how well such a model might simulate large-scale sea ice<br />

response. We also expect to gain insight into the essential features <strong>of</strong> the<br />

numerical results.<br />

In the model developed for <strong>AIDJEX</strong>, the yield surface depends on a<br />

strength parameter which is defined in terms <strong>of</strong> the thickness distribution.<br />

In order to simplify the present analysis as much as possible, we have<br />

eliminated this difficult nonlinearity and replaced it with the assumption<br />

that the material hardens slowly enough that an ideal plastic material model<br />

may be assumed. However, we allow spatial variations in the strength <strong>of</strong> the<br />

ice cover, but do not allow variations in this quantity with deformations.<br />

The yield surface is given by<br />

where we have expressed the yield surface in terms <strong>of</strong> the stress invariants<br />

G =<br />

I<br />

0 + G<br />

xx<br />

2<br />

and the yield strength p*, which is assumed to be given as a function <strong>of</strong><br />

position p* = p*(x,y). Furthermore, we find it convenient to rewrite ( 4)<br />

as<br />

which is shown schematically in Figure 1. From (6) we can find 011 whenever<br />

01 is known (since p* is given), and when the stress state is on the yield<br />

surface.<br />

The plastic flow rule presented by Coon et al. (1974) assumes that<br />

deformation is normal to the yield surface when flow occurs. In the present<br />

work it is to our advantage to generalize this flow rule to consider an<br />

arbitrary potential function:<br />

118


Fig. 1. Yield surface expressed in terms <strong>of</strong> invariants a1 and<br />

aII. Slope aoII/3o1 = b!(o,, p") is measured by angle 6 so<br />

that b' = tan B. The stretching vector lies at angle 8,<br />

which satisfies 8 - B = 7r/2 when a normal flow rule is<br />

assumed.<br />

If the potential function 11, is set equal to the yield function Cp in this<br />

expression, then the normal flow rule is obtained. Rewriting the flow rule<br />

in terms <strong>of</strong> the stress invariants gives<br />

where<br />

11, = ll,(uI, aII, p*) is a potential function<br />

o' - = o - - oI & is the stress deviator, and<br />

X 2 0 is a non-negative scalar.<br />

The potential function 11,<br />

is defined independently to describe the response<br />

<strong>of</strong> the material. If we choose I# to be <strong>of</strong> the form<br />

119


then the flow rule becomes<br />

Q = +A<br />

-<br />

--3’( 1 + --<br />

-<br />

.“I (10)<br />

I1<br />

The stretching is shown geometrically in Figure 1.<br />

given by<br />

The orientation <strong>of</strong> Q is<br />

where<br />

and<br />

The flow rule is chosen so that plastic flow is orthogoilal to the potential<br />

function $.<br />

Therefore, we have<br />

tan (0-.rr/2) = B’ (13)<br />

where 3’ = aB/ aG,.<br />

The normal flow rule<br />

Then B’ = tan f3 and 8-?r/2 = B-<br />

= $) provides that B = b I .<br />

CHARACTERISTIC ANALYSIS<br />

In the classical cases, plasticity analysis has been ccncerned primarily<br />

with determining loads at which deformations may occur. In these limiting<br />

cases it has been reasonable to neglect all motions and concentrate on the<br />

stress fields directly. Since our interest is in determining the motions<br />

described by the ice model, we cannot make such a simplifying assumption.<br />

However, the techniques used are similar. Our analysis follows Sch<strong>of</strong>ield<br />

and Wroth (P94&), who were interested In material models for soils. Although<br />

our material model is different, the choice <strong>of</strong> variables is the same.<br />

120


liv cnusidcring momtiitum balance., tlii> yic~ltl constr'iint, 'irid tlicl 1 low<br />

rule,, wc have six cquntions.<br />

the s ix dependent vari,ibLes '. 17. (1<br />

strt.ss invariants using (5).<br />

rliesc cquat ions may be exprc~ssc~d in tc>rnis <strong>of</strong><br />

llowevcar, to follow the rinalysis <strong>of</strong> Scli<strong>of</strong> itxld<br />

and "loth (1968) we eliminate Cartesian strcss components in favor <strong>of</strong> invari-<br />

dnts and. furthermore, eliminate maximum shear strcss OIT<br />

by substituticm<br />

using the yield constraint (4). The Cartesian stress components are related<br />

to invariants by<br />

xx' ('x\7*<br />

I1<br />

yy'<br />

'ind ,\ by i.1 imin'it ing thc<br />

o = CT sin 2y<br />

xy I1<br />

(5 = (5 - (T cos 2y<br />

YY 1 11<br />

where y is measured counterclockwise relative to the x-axis.<br />

Tiefore rewriting stress divergence in momentum balance in terms <strong>of</strong><br />

invariants, we eliminate the extra variable X from the flow rule. In<br />

Cartesian components. equation (10) hecoiiics<br />

But taking the ratio <strong>of</strong> (17) to (18) and (18) to (19) eliminates A+ As a<br />

result, we find two equations in u. v, oXx, oxy. and 0 . The equations are<br />

YY<br />

written as<br />

121


[-B' -F<br />

Finally, we eliminate Cartesian stress components from the momentum<br />

equations (3) and the flow rule (20-21) by substituting the equations (14-16).<br />

At the same time, we eliminate the maximum shear stress UII using the yield<br />

constraint (6). As a result, we obtain four equations in the four unknowns<br />

u, v, UI, and y:<br />

+(l+b'cos2y)--2bsin2y~+b'sin2y-+2bcos<br />

aoI aaI 2y3<br />

ax ax aY aY<br />

-<br />

-Tax - T~ (u-ug, v-vg) - mfc ( v-vg) --;;*cos 2y 32- ab sin 2y *<br />

ax ap* a3<br />

ab * ab a *<br />

= -T - T (u-ug,v-v +mfc (u-u,) -- sin 2y - -cos 2y-E<br />

ax x g aP * ax ap* aY<br />

It is useful to rewrite the system <strong>of</strong> equations in a matrix notation by<br />

affixing the four dependent variables into a solution vector 5:<br />

Then in matrix form the governing equations become<br />

where<br />

'$2 +L?g<br />

"X Y<br />

+ F = Q<br />

122


-rm 1 0<br />

I+]) ' cos 2y<br />

-271 sin 2y<br />

/I =<br />

.,<br />

0 -mzi<br />

2 sin 2y R' - cos 2y<br />

0 R' + cos 2y<br />

11' sin 2y 2h cos 2y<br />

0 0<br />

0 0<br />

-w 0<br />

0 -mu<br />

B' - COS 2y 0<br />

0<br />

0<br />

(29)<br />

0<br />

0<br />

and the inhomogeneous contribution is<br />

F - =<br />

0<br />

0<br />

Both the coefficients 4 and and the forcing function < depend on the solution,<br />

but not derivatives. The system is quasi-linear and we can find<br />

characteristics by considering the system locally. It is interesting to<br />

note that air stress, water stress, Coriolis acceleration, and spatial variability<br />

<strong>of</strong> p* do not affect the principal part <strong>of</strong> the governing equations,<br />

but enter only the driving force on the right-hand side. This implies that<br />

none <strong>of</strong> these forces can affect the existence or orientation <strong>of</strong> characteristics<br />

(except that the driving forces affect the solutions, <strong>of</strong> course).<br />

Characteristic curves have the following properties, each <strong>of</strong> which can<br />

be used as a definition (Courant and Hilbert, 1962):<br />

1. Along a characteristic curve the differential equation (or, for<br />

systems, a linear combination <strong>of</strong> the equations) represents an interior<br />

123


differential equation.<br />

2. Discontinuities <strong>of</strong> a solution cannot occur except along characteristics.<br />

3. Characteristics are the only possible branch lines <strong>of</strong> solutions,<br />

i.e., lines for which the same initial value problem may have several solutions.<br />

The first property, which shall be exploited in the derivation <strong>of</strong> characteristic<br />

directions and relations, is inherent in the basic fact: A direction<br />

is characteristic at a point P if there exists a linear combination <strong>of</strong> the<br />

differential equations for which all the unknowns are differentiated at P only<br />

in this direction. A system <strong>of</strong> equations is said to be hyperbolic if it can<br />

be replaced by a linearly equivalent one in which each differential equation<br />

contains at every point differentiation in only one characteristic direction.<br />

The second and third properties will be more useful in interpreting the<br />

results <strong>of</strong> this analysis.<br />

In this section we present formally the approach both for determining<br />

the existence <strong>of</strong> characteristic directions and for determining the equations<br />

that hold along each such curve.<br />

Assume we are given the system <strong>of</strong> n equations in n anknowns:<br />

which is chosen to have an appearance identical to the system <strong>of</strong> the four<br />

equations shown in (27). If we pre-multiply by the vector RT to obtain<br />

R T (43, + BZ + E) = 0,<br />

.<br />

--Y<br />

we obtain one single scalar equation which is a linear combination <strong>of</strong> the<br />

equations.<br />

If there exist n distinct roots and n eigenvectors to the system<br />

RL (Aa A, - E ) = 0<br />

-a<br />

a = 1, 2, ..., n<br />

(33)<br />

then for each such eigenvector R<br />

'a<br />

that<br />

there is a linear combination <strong>of</strong> (31) so<br />

124


m I-<br />

a = 1, 2, ..., il<br />

(34)<br />

which is a set <strong>of</strong> n equations in z.<br />

If we consider the curve < shown in Figure 2 and parameterize it<br />

in terms <strong>of</strong> 5 when x = x(


Substituting (36) and (37) into (34) provides one equation for each<br />

characteristic direction. It is<br />

CY,<br />

where 5<br />

is the coordinate associated with X a'<br />

In summary, the characteristic directions and governing equations are<br />

obtained by the following sequence. From the equation (33) determine the<br />

n distinct roots A which define the characteristic directions K These<br />

a<br />

a'<br />

roots arise when the coefficient matrix is singular:<br />

With the roots thus determined, the eigenvectors R are found from (33).<br />

-a<br />

Since the governing equation (32) is homogeneous, the magnitude <strong>of</strong> R is<br />

-a<br />

immaterial. Finally, along each characteristic direction the solution vector<br />

satisfies an ordinary differential equation (38). In theory, the n equations<br />

may then be integrated, each along its appropriate direction.<br />

EXISTENCE OF CHARACTERISTIC CURVES<br />

For our system <strong>of</strong> four equations in four unknowns we consider the eigenvalue<br />

problem given by equation (33). Our task is to find four values <strong>of</strong> X<br />

a<br />

such that the corresponding non-zero R may be found. It is a well-known<br />

-a<br />

property <strong>of</strong> linear algebraic systems <strong>of</strong> equations that the coefficient matrix<br />

(XaA-€l)<br />

must be singular. Therefore, we set<br />

det (Aa 4-B) = 0<br />

(39 bis)<br />

and seek roots X<br />

a<br />

substituted into this equation from equations (28) and (29).<br />

that satisfy the equation. The elements <strong>of</strong> 4 and B may be<br />

Expansion <strong>of</strong><br />

the determinant and determination <strong>of</strong> the roots from the characteristic polynomial<br />

could then follow. However, the special structure <strong>of</strong> the matrix A4-B<br />

allows a simpler approach. We note that after substitution we may write<br />

126


where 1 is the two-by-two identity matrix, 0 is the two-by-two zero matrix<br />

and<br />

c= (41)<br />

c=<br />

Expansion <strong>of</strong> the determinant by c<strong>of</strong>actors leads to<br />

det (A A_-e) = det c det C<br />

(43)<br />

Therefore, the roots <strong>of</strong> equation (39) may be found from<br />

det = 0<br />

(44)<br />

and<br />

det C = 0 (45)<br />

It is interesting to note that advection has no effect on the characteristic<br />

directions. The existence and direction <strong>of</strong> characteristic curves<br />

depend only on properties <strong>of</strong> the yield curve and flow rule. Perhaps it is<br />

not surprising that the water stress, Coriolis force, and other applied body<br />

forces have no effect; but since the velocity gradient components appear in<br />

the advection terms, it seems surprising that characteristic curves may be<br />

found without considering this term.<br />

The elements <strong>of</strong> the matrix arise from the stress divergence term in<br />

the momentum balance equations. Therefore, we call the directions obtained<br />

by satisfying equation (44) the stress characteristics. Also, we<br />

12 7


arbitrarily assign the values <strong>of</strong> 1 and 2 to index a for these two directions.<br />

Similarly, the elements <strong>of</strong> arise in the flow rule and involve velocity components.<br />

Therefore, we call the directions obtained from equation (45) the<br />

velocity characteristics and assign to index a the values 3 and 4 for these<br />

roots.<br />

Equation (44) is expanded by replacing X by K<br />

a a<br />

gent by sine and cosine. It becomes<br />

and expressing the tan-<br />

I[-(l+b' cos 2y) sin K - b' sin 2y cos K ][2b cos 2y sin K~ - 2b sin 2y cos K ~ ]<br />

a<br />

a<br />

-[b ' sin 2y sin K - (1-b ' cos 2y) cos K ][-2b sin 2y sin K - 2b cos 2y cos K ] = o<br />

a a a a<br />

which reduces to<br />

(46)<br />

(b' +cos 2y) sin2 K ~ -<br />

2 sin 2y sin K<br />

a cos K<br />

a + (b' - cos 2y) cos2 K a = o (47)<br />

Rut substituting the double angle formulas for sin2 K a' cos2 K a and sir, K a cos K a<br />

provides<br />

(b'+cos 2y) (1-cos 2 ~ - ~ 2sin ) 2ysin 2~ + (b'- cos2y)(1+cos 2~ a ) = 0 (48)<br />

a<br />

which simplifies to<br />

b' - cos 2y cos 2~ - sin 2y sin 2~ = 0 (49)<br />

a<br />

a<br />

Finally, using the trigonometric expansion for the cosine <strong>of</strong> the difference<br />

<strong>of</strong> two angles gives<br />

1) , then<br />

If the slope <strong>of</strong> the yield curve b' is replaced by the angle f3<br />

(see Fig.<br />

tan P = cos 2(y - K ~ ) a = l,2 (51)<br />

relates all the angles that define the stress characteristic directions.<br />

Similarly, equation (45) may be expanded to the form<br />

12 8


[2 sin 2y sin K~ - - cos 2y) cos K I[ (;: ' + cos 2y) sin K - 2 sin 2y cos K ]<br />

c1 a a<br />

This expression reduces to<br />

- [ - (P' +cos 2 ~ cos ) K,][(B'<br />

- cos 2y) sin K 0. 3 = o<br />

(52)<br />

(B' +cos 2y) sin' K - 2 sin 2y sin K cos K + (B' - cos 2y) cos2 K = o (53)<br />

a a a a<br />

which is identical to equation (47) with b' replaced by B'.<br />

find that<br />

Therefore, we<br />

The velocity characteristics are thus defined by a relationship similar to<br />

that for the stress characteristics. If we substitute for B' in terms <strong>of</strong> the<br />

angle 8 which relates shearing to dilating, then from equation (13)<br />

tan (8 - n/2) = cos 2(y - K ) a= 3,4 (55)<br />

a<br />

These transcendental equations will be useful for interpreting the exis-<br />

tence <strong>of</strong> characteristic directions. However, to transform the governing<br />

equations into their appropriate forms along the characteristic coordinates<br />

requires that we have an explicit expression for X . These expressions may<br />

a<br />

be found by considering equations (47) and (53) directly, substituting h<br />

a<br />

for tan K . The roots <strong>of</strong> the quadratic equations then are<br />

a<br />

a<br />

= I-<br />

sin 2y+ J 1 - (b')'<br />

b'-+--cos 2Y<br />

a = 1,2<br />

and<br />

sin 2y d 1 - (~')2<br />

a I?' + cos 2Y<br />

A =<br />

a = 3,4 (57)<br />

It is seen that when a non-normal flow rule is assumed, four distinct characteristic<br />

roots and directions occur. However, when a normal flow rule is<br />

assumed, two pairs <strong>of</strong> repeated roots arise and the stress characteristics<br />

coincide with the velocity characteristics. In this latter case, the governing<br />

equations along characteristic directions cannot be obtained from<br />

equation (38) when advection is included because the formal result breaks down.<br />

129


The special case when advection can be neglected is considered in a later<br />

sect ion.<br />

In Figure 3 we show graphically the relationship between B (or 8 - n/2)<br />

and K - y. The characteristic direction K appears as the abscissa and yield<br />

curve slope 8 as the ordinate. This implies that, to use the graph, we will<br />

be given K and y and are to find 8.<br />

However, in our analysis it is K that<br />

we seek to determine and that variable depends on 8 and y. The analysis<br />

would dictate that we plot 8 as the abscissa and K as the ordinate. Our<br />

desire to exchange axes is brought about by looking ahead to the use <strong>of</strong> data<br />

to help evaluate the ice model. In that application we anticipate looking<br />

for a relationship between lead patterns and characteristic directions, and<br />

these are the data that will be given. The results presented in Figure 3<br />

then will help us to determine the shape <strong>of</strong> the yield curve.<br />

cs<br />

z<br />

I35<br />

I20<br />

IO5<br />

90<br />

75<br />

60<br />

45<br />

c.r<br />

v)<br />

a,<br />

?! I<br />

L<br />

I<br />

I<br />

I<br />

tanP = cos2 I K - y I<br />

\<br />

3U<br />

I \<br />

Fig. 3. Relationship between slope <strong>of</strong> yield curve B and characteristic<br />

direction K relative to maximum principal stress<br />

direction y.<br />

Several comments must be made about the results given by (51). Similar<br />

comments are relevant for (55). The characteristic direction is given by<br />

130


lines through each point.<br />

ways to traverse each line.<br />

period TT. This fact is reflected in the expressions. We see that real char-<br />

acteristic roots arise from (56) whenever IF’] - < 1. Thus, two distinct<br />

directions exist except at the upper limit.<br />

directions exist, the system <strong>of</strong> equations is hyperbolic.<br />

are identical, the system <strong>of</strong> equations is parabolic.<br />

When two distinct characteristic<br />

Where the two roots<br />

Finally. whenever<br />

Ih’l > 1, no real roots may exist and the system <strong>of</strong> equations is elliptic.<br />

As stated in the introduction, these different characterizations <strong>of</strong> the<br />

equations are important primarily because discontinuities in the solutions<br />

can exist only across real characteristic lines.<br />

We cannot distinguish between the two possible<br />

Therefore, any result must be repetitive over<br />

To help visualize the patterns <strong>of</strong> characteristic curves that may appear,<br />

the results <strong>of</strong> equations (50) and (55) are displayed in Table 1 and graphi-<br />

cally in Figures 4 and 5.<br />

In each <strong>of</strong> the six examples a homogeneous stress<br />

state is assumed and the x-axis (horizontal) is the direction <strong>of</strong> maximum<br />

principal stress (y = 0).<br />

Since the plastic response is independent <strong>of</strong> the<br />

Fig. 4. Stress invariants and stretching vectors associated with six<br />

sample conditions: (a) isotropic contracting, (b) uniaxial contracting,<br />

(c) shearing and closing, (d) pure shearing, (e) shearing<br />

and opening, (f) uniaxial extending.<br />

131


TABLE 1<br />

STRETCHING, STRESS, ASP CHARACTERISTIC DIRECTION PARAMETERS FOR SIMPLE<br />

HOMOGENEOUS FIELDS<br />

-1 0 -112 -112 0 n -1.00 -1.00 n/2 none Sa<br />

-&I2 &I2 0 4 1 2 0 3n/4 -0.67 -1.11 nI4 0 5b<br />

-112 h/2 $-l+ 5) $-1-m 0 2n/3 -0.55 -1.10 n/6 t27.4 5c<br />

0, 1 112 -112 0 nI2 -0.26 -0.93 0 f45.0 5d<br />

112 hl2 +A) +cl-A, 0 n/3 0.00 -0.44 -n/6 262.6 5e<br />

m 2 &I2 JT/2 0 0 714 0.04 -0.20 -n14 90 5f<br />

132


-<br />

W<br />

L<br />

-<br />

I-<br />

0<br />

a<br />

e<br />

I-<br />

Z<br />

0<br />

u<br />

u<br />

-'<br />

a<br />

0<br />

a<br />

I-<br />

O<br />

v,<br />

-<br />

1 1 1 1 1<br />

A<br />

0<br />

Y<br />

133


ate <strong>of</strong> deformation, we have normalized p such that D; + D:=<br />

= 1 and each <strong>of</strong><br />

the six examples then corresponds to a different ratio <strong>of</strong> shearing to dilating,<br />

measured by 8. The stretching invariants are presented in the first two<br />

columns <strong>of</strong> Table 1. The normalized principal values to be used in Figure 5<br />

appear in columns 3 and 4; and y, the orientation <strong>of</strong> p, appears in column 5.<br />

The angle 8 is given in column 6. For the yield surface presented in Figure<br />

5 (we have chosen an arbitrary but representative curve) the stress states<br />

needed to induce each <strong>of</strong> the stretchings are shown. These principal values<br />

are given in column 7 and 8 <strong>of</strong> Table 1. For the normal flow rule the princi-<br />

pal directions <strong>of</strong> D and o are aligned. The tangent to the yield curve is<br />

., .,<br />

given in column 9. Orientation <strong>of</strong> the characteristic directions is described<br />

in column 10. Finally, in column 11, we state which part <strong>of</strong> Figure 5 gives<br />

the proper graphic description.<br />

Results are understood most simply if the reader assumes that a normal<br />

flow rule is applicable SO that stress and velocity characteristics coincide.<br />

However, for the more general non-normal flow rule, the results describe<br />

either the stress characteristics or the velocity characteristics, but not<br />

both. To visualize the stress characteristics, select the example according<br />

to the value <strong>of</strong> 6 associated with either the stress or the stretching state.<br />

For a specific flow rule, 8 and (3 are known for each stress state so that<br />

both pairs <strong>of</strong> characteristics may be superposed to visualize the complete<br />

set <strong>of</strong> characteristic directions.<br />

In Figure 5a isotropic contracting is presented. For this case no real<br />

characteristic directions exist because 8 > 3~14. Uniaxial contracting is<br />

shown in Figure 5b, One real characteristic direction exists and it is<br />

oriented parallel to the directions <strong>of</strong> the larger principal stress. For this<br />

case we could envision that ridges form along the characteristic directions<br />

to allow the deformation to occur. The ridges are orthogonal to the direction<br />

<strong>of</strong> maximum normal stress. We must remember that this model does not<br />

describe the fornation <strong>of</strong> a single ridge which must form along an existing<br />

lead, but instead represents the average <strong>of</strong> an isotropic field <strong>of</strong> ridges.<br />

However, it is doubtful that the analysis is valid for this condition in<br />

any case. When ridging occurs we expect a hardening model rather than an<br />

ideal plastic model to be required. A hardening plastic model has not been<br />

134


considered and it might alter the results, perhaps by eliminating all charac-<br />

teristic directions. Thus, we do not consider seriously the cases where<br />

0 n/2 for which DI < 0. We do note that for the assumed yield surface<br />

(Figure 4) we require a confining stress to obtain uniaxial contracting.<br />

In Figure 5c a combined stretching <strong>of</strong> shearing and closing is presented.<br />

In this case, the shearing exceeds the closing and DII/DI = -6. The stress<br />

state is similar to that which induces uniaxial contracting, but the differ-<br />

ences are large enough to change the characteristics. For 8 = 2~13 we have<br />

two real characteristic directions oriented symmetrically about principal<br />

directions at angles k27.4 degrees.<br />

In Figure 5d the case <strong>of</strong> pure shearing is sketched. Since no area<br />

changes occur we have 8 = T/2 and two characteristic directions <strong>of</strong> ?45<br />

degrees.<br />

This is the classical result that occurs with a von Mises yield<br />

surface. In our case it occurs only at the point <strong>of</strong> maximum shear stress.<br />

We can imagine velocity discontinuities across these lines associated with<br />

shearing across leads.<br />

In Figure 5e we present the case <strong>of</strong> shearing and opening. We see that<br />

for the particular yield surface chosen this stretching requires a uniaxial<br />

stress state. The stretching is similar to that in Figure 5c but the sign<br />

<strong>of</strong> DI has been changed from negative to positive.<br />

also similar but the pressure 01 is different.<br />

The stress deviator is<br />

We again have two character-<br />

istic directions but in this case they are oriented at 262.6 degrees from<br />

the larger principal stretching. Thus, although this characteristic pattern<br />

could be obtained by rotating the material 90 degrees as in Figure 5c, the<br />

stress and stretching are not the same. Therefore, we must exercise care<br />

when attempting to determine what state exists. Again, in this case we can<br />

visualize leads that are shearing and opening along characteristic directions.<br />

Finally, in Figure 5f we present uniaxial opening. For the yield sur-<br />

face presented the material may support a small amount <strong>of</strong> tensile stress,<br />

but there is nothing critical about this property. One real characteristic<br />

occurs at 90 degrees, which is orthogonal to the direction <strong>of</strong> maximum open-<br />

ing.<br />

It is easy to visualize leads opening along this characteristic<br />

direction for the uniaxial opening case.<br />

135


CHARACTERISTIC EQUATIONS<br />

The ordinary differential equations that must be satisfied along each<br />

real characteristic direction have been discussed formally in an earlier section<br />

(eqs. 31-39). In this section we look at several special cases that<br />

help us to understand the general result, and we discuss the complete set <strong>of</strong><br />

equations e<br />

Eigenvectors 2-<br />

-a<br />

<strong>of</strong> equation (33) must be found for each characteristic<br />

direction A These vectors can then be used in equation (38) to derive the<br />

a'<br />

gcverning characteristic equations. If we divide the coefficient matrix<br />

E) into appropriate 2x2 matrices (40). then it is desirable also to<br />

4-<br />

divide the eigenvectors similarly, To this end, let<br />

R =<br />

-a<br />

S<br />

a = 1, 2, 3 , 4<br />

where i7<br />

-a<br />

&?,(Aa 4 - @><br />

-<br />

and Sa are each two<br />

and express the<br />

the eigenvectors satisfy the<br />

component vectors. If we form the product<br />

result in terms <strong>of</strong> the submatrices, we find that<br />

relationships (for each a)<br />

m<br />

m<br />

After finding R and S from this linear algebraic set <strong>of</strong> equations, we<br />

-a -a! m<br />

substitute into the expression Q' A_, whichmay be rewritten as<br />

-a<br />

where<br />

Ai2 Y =<br />

i<br />

b' sin 2y 2b cos 2y<br />

(2 sin 2y B' - COS 21,<br />

B' + cos 21,<br />

136


Now also divide the solution vector and forcing function into<br />

and<br />

where<br />

= [i], C_ = [y], and F1 = r"] contains the first ti 3 compon-nts <strong>of</strong><br />

- \f2<br />

Substituting these new variables and expanding the matrix products trans-<br />

E.<br />

forms the governing equation (38) into<br />

This expression showsdirectly how the velocity and stress interact along<br />

each <strong>of</strong> the four characteristic curves and the influence <strong>of</strong> advection and<br />

the driving force.<br />

vectors R<br />

We consider equations (59) and (60) to determine explicitly the eigen-<br />

-01<br />

(alternately Ga and 8,).<br />

Several general properties <strong>of</strong> the<br />

eigenvectors R associated with the stress characteristic (a = 1,2 and<br />

-a<br />

det ca = 0) may be identified. Since c is singular we use (60) to find<br />

-a<br />

R that are nonzero. Then, if all eigenvalues are distinct, C is not singu-<br />

-a -a<br />

lar, and so the nonhomogeneous albegraic equations (59) may be solved for<br />

S . A comparable consideration <strong>of</strong> the UeZocity characteristics (a = 3,4 and<br />

-a<br />

det c, = 0) requires that (60) give G, = 0 because c is nonsingdar. Using<br />

(59) and the singularity <strong>of</strong> C provides nonzero S from the homogeneous<br />

-a -a<br />

equations.<br />

For materials for which a normal flow rule is assumed, the stress and<br />

velocity characteristics coincide and both C and c are singular simultane-<br />

-a -a<br />

ously. From (60) we find a nonzero vector R But then in (59) we find a<br />

-a'<br />

nonhomogeneous system <strong>of</strong> equations with a singular coefficient matrix C<br />

-a'<br />

While it is possible that values <strong>of</strong> R and S exist for this case, special<br />

-a -a<br />

care must be exercised to determine the eigenvectors.<br />

*<br />

-01<br />

13 7


One way to circumvent the problems that arise when stress and velocity<br />

characteristics coincide is to neglect advection. In this classical approach,<br />

the stress and velocity are found by considering the two systems <strong>of</strong> equations<br />

whose principal parts are uncoupled.<br />

Systems Uncoupl ed by Neglecting Advection<br />

If we consider the special case in which advection is neglected, the<br />

results take a simpler form. The equations governing stress and the equations<br />

governing velocity uncouple in their first derivatives so that if tractions<br />

are presecribed everywhere along the boundary, and velocity-dependent body<br />

forces are zero, a statically determinant problem is defined. In this case<br />

the stress state may be found independently <strong>of</strong> the velocity field.<br />

The equations (59) and (60) for finding JLa<br />

reduce to<br />

T<br />

s c = g<br />

-a -a<br />

If we seek the stress characteristics given by a = 1,2 then<br />

det 7 = 0 det -C, # 0 (69)<br />

-a<br />

and from the latterpropertywerequire thatS = fora=land2. On the other<br />

a<br />

hand, since ,C is singular we can find nonzero vectors e Thus, let<br />

3 0-<br />

and then find q and r from equation (68) expanded into<br />

a a<br />

components<br />

(9, Fa)<br />

Xa(l+b' cos2y) -b'si~2y -2X bsin2y- 2b cos 2y<br />

a<br />

X b ' sin 2y - (1 - b ' cos 2y) 2Xa b cos 2y -<br />

2b sin 2y<br />

= (0 0) (71)<br />

If we arbitrarily choose the length <strong>of</strong> vectors aa by letting<br />

r2 = 1 then it can be shown (after some algebra) that<br />

= 1 and<br />

138


q1 = -A2<br />

q2 =<br />

A1<br />

Therefore, the vectors<br />

and e2 are<br />

R2 = [11]<br />

(74)<br />

Making use <strong>of</strong> the zero terms in coefficients, equations along stress characteristics<br />

are found by substituting for R and sa, a = 1,2 into equation<br />

-a<br />

(66) neglecting the advection. They become<br />

A similar analysis <strong>of</strong> equations along a ueZocity characteristic given<br />

by a = 3,4 requires that eigenvectors<br />

-a<br />

R<br />

and (68) with<br />

Then, since ca'is nonsingular, we require that<br />

be determined from equations (67)<br />

det ca = 0 det ?& # 0 (76)<br />

R = O<br />

-a -<br />

(77)<br />

But S can be found from (67) by solving the linear algebraic equations. It<br />

-a<br />

can be shown that<br />

where the first component has arbitrarily been chosen to be unity. Again<br />

making use <strong>of</strong> zero coefficients and neglecting advection, the governing<br />

equations (66) become<br />

139


The equations governing solutions along stress and velocity characteristics<br />

(75) and (79) can now be written explicitly in terms <strong>of</strong> the components<br />

u, u, OI and y. However, as we determine the specific form taken by each<br />

characteristic equation, the use <strong>of</strong> a subscript ci = 1, ..., 4 becomes cumbersome.<br />

Therefore, it is helpful to introduce another notational scheme to be<br />

used to express final results in term <strong>of</strong> components. In the new notation,<br />

eigenvalues associated with stress characteristics are defined by<br />

The orientation angles associated with these roots are IC+<br />

-<br />

where<br />

tan K+ =<br />

-<br />

Ot<br />

and the coordinates along each stress characteristic curve are<br />

A similar set is introduced to define velocity characteristics. Characteris-<br />

tic roots are<br />

?J, = A3<br />

4<br />

The associated angles v,<br />

-<br />

satisfy<br />

tan v, =<br />

-<br />

1-1,<br />

(84)<br />

and the coordinates along each velocity characteristic curve are<br />

The use <strong>of</strong> the symbol f is especially useful because roots and equations<br />

along either stress or velocity characteristics differ only in a few signs,<br />

If we return to the governing equation along stress characteristics


where the roots are<br />

sin 2y<br />

- b’ + cos 2y<br />

0, =<br />

and the cosine may be found from trigonometric identities and equation (81)<br />

as<br />

The direction <strong>of</strong> increasing values <strong>of</strong> c1 and S2 are then required to be con-<br />

5 K<br />

sistent with -~/2 < ~/2, or positive x. The coefficients <strong>of</strong> equation<br />

a-<br />

(86) may be simplified somewhat. It can be shown that<br />

The cosine becomes<br />

and, when divided by the bracketed portion <strong>of</strong> the second coefficient in<br />

equation (86), may be written as<br />

cos K+_<br />

-<br />

b’ + cos 2y<br />

OF sin 2y + cos 2y<br />

JZ [l+b cos 2y sin 2y 13<br />

(92)<br />

Finally, substituting all appropriate terms into the governing equations<br />

(86) provides<br />

141


It is readily seen that when body forces are nonzero ($1 # 0, $2 # 0) the<br />

equations are far more complicated than when body forces are absent. For<br />

the statically determinant case we would probably assume that the material<br />

is at rest so that only the applied air stress is contained in f1 and $2.<br />

However, if we also consider a nonzero velocity field, then $1 and $2 contain<br />

contributions from water drag, Coriolis force, and strength gradients given<br />

by equation (30).<br />

We now return to the governing equations along velocity characteristics<br />

(79), and follow an analogous development. Expansion into components and<br />

use <strong>of</strong> the new notation provide two simple governing equations:<br />

where the roots are<br />

- sin 2y i d i g<br />

-<br />

1-11 B’ + cos 2y<br />

(94)<br />

These equations are seen to be homogeneous and coefficeints are less complex<br />

than the stress characteristic equations. However, the directions 1-1, depend<br />

on the stress state through both y and B’ and so the difficulties <strong>of</strong> determining<br />

stress cannot be eliminated.<br />

An important physical property <strong>of</strong> the velocity field is that no stretching<br />

occurs along velocity characteristic curves. To demonstrate this property<br />

for the <strong>AIDJEX</strong> plastic model shown for classical plastic models by Hill<br />

(1950), Martin (1975), and others, consider the requirement that at a point<br />

there be no stretching in an arbitrarily chosen direction given by unit<br />

vector<br />

-<br />

where @ indicates the counterclockwise angle between n and the x-axis. If<br />

there is no stretching along this direction, n remains a unit vector and<br />

142


The material time rate <strong>of</strong> change <strong>of</strong> equation (96) provides<br />

m<br />

(97)<br />

where we have used the fact that the material time value <strong>of</strong> change <strong>of</strong> 5 is<br />

.<br />

n = & g (98)<br />

In components, equation (97) becomes<br />

2<br />

D cos 4 + 2Dxr cos 4 sin 4 + D sin2 4 = 0<br />

xx<br />

YY<br />

(99)<br />

The Cartesian components <strong>of</strong> the stretching tensor may be expressed in<br />

terms <strong>of</strong> invariants DI and D and the principal direction y [see equations<br />

II<br />

(14-16) for similar expressions for the stress tensor]:<br />

2<br />

D =--- D1 D1l cos 2y<br />

YY 2 2<br />

Substituting these expressions into equation (99) and using the formula for<br />

the cosine <strong>of</strong> the difference <strong>of</strong> two angles provides<br />

DI + DII COS 2(@ - y) = 0<br />

But the ratio <strong>of</strong> shearing to dilating is<br />

-- DII - tan 0<br />

DI<br />

(11 bis)<br />

Rearranging to find the inverse provides<br />

tan(0 - ~ /2)<br />

DI<br />

= - -<br />

DII<br />

143


which may be substituted into equation (103) to obtain<br />

This analysis shows that at each point there are two orientations @<br />

that satisfy equation (105) and along which there is no stretching. A comparison<br />

<strong>of</strong> this result with equation (55) then provides the fact that$<br />

corresponds to the directions <strong>of</strong> the velocity characteristics (K a = 3,4)<br />

a'<br />

and therefore that no stretching occurs along velocity characteristic curves.<br />

The analysis, however, has not shown clearly what happens when characteristics<br />

coincide or do not exist: 0 <<br />

- IT/^ or 0 ><br />

-<br />

3~14. These special<br />

cases are seen easily by considering the Mohr's circle for Q shown in Figure<br />

6. (Thefollowing approach was pointed out to the authors by J. F. Nye.)<br />

Fig. 6. Mohr's circle for stretching<br />

tensor 0.<br />

The construction <strong>of</strong> Mohr's circle follows standard procedures (e.g.,<br />

Grandall and DahL, 1957). The states a and b in Figure 6 represent stretching<br />

expressed in x,y coordinates. The points c and c' represent directions<br />

144


in which there is shearing but no extending since these lie at the origin<br />

<strong>of</strong> the horizontal axis that is a measure <strong>of</strong>,normal stretching.<br />

is shown as the angle from the x-axis to point c and appears as 2Cp in the<br />

Mohr's circle.<br />

axis and are at +(Cp-y).<br />

The angle Cp<br />

Points c and c' are symmetric about the principal stretching<br />

Geometrically, it is clear that when IDI I > DII<br />

the circle does not intersect the origin, and so there cannot be directions<br />

<strong>of</strong> no stretching. It is also clear that when lDIl = DII the two roots for<br />

4 coincide providing only one direction <strong>of</strong> no stretching. These special<br />

cases occur when the equations are elliptic and parabolic, respectively.<br />

The General Non-Normal Flow Rule Case<br />

When advection is included in the momentum equation, the eigenvectors<br />

are somewhat different from the simplest uncoupled system. Our attention<br />

returns to equations (59) and (60) to define the eigenvectors R a = 1, 2,<br />

-a<br />

We note first that the eigenvectors associated with the veZocity<br />

..., 4.<br />

characteristics (a = 3,4) are unchanged by including advection. These are<br />

defined by det C = 0 and for distinct roots (non-normal flow rule) require<br />

-a<br />

that det c 0. As before, we obtain R = 9. Then S satisfies (59) or<br />

-a -a -a<br />

(67) and is given by (78). The equation governing the solution along these<br />

characteristics remains the same (93) so that the velocity components still<br />

satisfy the simple relationships.<br />

as given in (94).<br />

The roots are also unchanged and remain<br />

A similar consideration <strong>of</strong> the equations governing solutions along<br />

stress characteristics shows changes. The stress characteristics (a = 1,2)<br />

are defined by det ?? = 0 and det # 0. Thus, from equation (60) we find<br />

-a<br />

the eigenvectors R to be unchanged from the uncoupled case and given by<br />

-a<br />

equations (73) and (74). But by (59) we find that S is no longer zero.<br />

-a<br />

Instead, we have<br />

since C is nonsingular. Substituting into equation (66) and evaluating<br />

--a<br />

all coefficients provide the equations governing solutions along these<br />

characteristics. Tt is seen that the forcing function and coefficient <strong>of</strong><br />

145


a<br />

the stress derivative dCldS remain as in equation (75). However, the veloc-<br />

-<br />

ity derivatives dUldSa also appear in the stress characteristic equation so<br />

that uncoupling does not occur. The coefficient <strong>of</strong> the velocity derivative<br />

dlJ/dSa - is<br />

T T T T<br />

-m u Ra + sa 421 = -m u R + m(X u-v) R C-' A21<br />

-a c1 -a -a -<br />

Some simplification <strong>of</strong> this coefficient occurs if we combine terms and sub-<br />

stitute Ca = Aa 421 - G21e<br />

Then the equation governing solutions along<br />

stress characteristics may be written as (a = 1,2)<br />

g21 =<br />

B? - COS 2y<br />

B' 1- COS 2y<br />

2 sin 2y<br />

It is clearly possible to express the above matrix equation in component<br />

form. However, we shall not write out the component equations, simply<br />

because the coefficients are quite lengthy and we do not intend to use the<br />

result explicitly. The above form <strong>of</strong> the equation is useful in our analysis,<br />

though, because it provides a concise description showing which terms appear.<br />

This general result is expected to be useful in future work when results for<br />

a normal flow rule are studied.<br />

Taking the Limit to the Normal Flow Rule<br />

The normal flow rule may be represented In the previous development by<br />

letting B' = br so that the potential function defining the direction <strong>of</strong><br />

flow 9 is identical to the yield surface 4. It is seen that the character-<br />

istic directions (given by ha, a = 1,2,3,4) are not distinct in this case.<br />

That is, the stress and velocity characteristics coincide (A3 = Al, A!, = X2).<br />

In this case ca and ca are singular simultaneously.<br />

The previous analysis<br />

breaks down when attempting to determine the eigenvectors associated with<br />

146


the stress characteristics, S a = 1,2. Since C is singular, equation<br />

-a ’ -a<br />

(106) is meaningless. However, our intuition leads us to believe that the<br />

model should not change dramatically when a normal flow rule is introduced.<br />

It also seems that the characteristic analysis should not become invalid.<br />

It is felt further that the governing equations possess finite limits. Our<br />

future efforts are aimed at determining these limits.<br />

CONCLUSION<br />

In the introduction to this paper five goals <strong>of</strong> this work were stated.<br />

Although no one goal has been met entirely, there has been substantial,<br />

progress toward all <strong>of</strong> them.<br />

Our understanding <strong>of</strong> discontinuities has improved. We have determined<br />

that the differential equations governing quasi-steady, rigid-plastic ice<br />

models cannot be simply characterized because theymaybelocallyhyperbolic,<br />

parabolic, or elliptic, depending on the state <strong>of</strong> stress (or stretching) at<br />

each point. The ratio <strong>of</strong> shearing to dilating ocntrols the character <strong>of</strong><br />

the system. If shearing predominates ( ~ / < 48 < 3?r/4), the system is hyperbolic<br />

and two distinct real characteristic directions exist, But when<br />

dilating is predominant (either opening, 8 < ~/4, or closing, 8 > 3~/4),<br />

the system is elliptic and no real characteristic directions exist. In the<br />

cases <strong>of</strong> uniaxial extension (e = ~/4) and uniaxial contraction (0 = 3~/4),<br />

the system is parabolic and one real characteristic direction occurs. The<br />

crucial point to be made is that the plasticity model can admit discontinuous<br />

solutions whenever real characteristic curves exist. This property is<br />

dramatically different from other models such as viscous models which are<br />

always elliptic during quasi-steady flow. We believe discontinuities are<br />

necessary to allow the explanation <strong>of</strong> such diverse features as spatially<br />

smooth velocity fields, shear ridging, and large leads formed within the<br />

pack ice. We know <strong>of</strong> no model other than a plastic model that allows a<br />

natural representation <strong>of</strong> all these features.<br />

It is anticipated that interpretation <strong>of</strong> numerical solutions will be<br />

enhanced by knowing the conditions under which discontinuities may appear.<br />

147


Although we have not studied in this work the magnitude <strong>of</strong> the discontinuities,<br />

there is information available elsewhere on this topic.<br />

The system <strong>of</strong> partial differential equations governing model response<br />

has been transformed into a system <strong>of</strong> ordinary differential equations with<br />

derivatives occurring only along characteristic curves. This analysis has<br />

been completed for two somewhat different modifications to the rigid-plastic<br />

AIDSEX model: (1) advection is neglected, and (2) the flow rule is modified<br />

to be non-normal to the yield surface. We have as yet been unable to analyze<br />

the AlDJEX model when advection is considered and a normal flow rule is<br />

assunled.<br />

The characteristic directions at each location depend on the stress or<br />

stretching states. The existence and orientation <strong>of</strong> the characteristic<br />

curves are independent <strong>of</strong> advection, air stress, water drag, Coriolis force,<br />

sea surface tilt, and yield strength gradients, except as the terms affect<br />

the stress state. Variations in stress along the stress characteristics,<br />

however, is affected by these terms. Along velocity characteristic curves<br />

there can be no stretching. Although this property does not provide a<br />

physical explanation for a correspondence between leads and characteristics,<br />

neither does it provide evidence that no such relationship exists. Inongoing<br />

work we intend to study directly whether or not there is a correlation<br />

between lead patterns and the stretching tensor. Preliminary work suggests<br />

that during unfaxial opening, leads form orthogonally to the direction <strong>of</strong><br />

maximum opening. This result, however, is easily anticipated. We have not<br />

yet found any correlation in other stretching states.<br />

If such a correlation between lead orientations and characteristics can<br />

be found, we intend to apply the theory to use satellite imagery to compare<br />

with characteristic directions calculated in two-dimensional simulations <strong>of</strong><br />

sea ice dynamics. This will provide a simple indication <strong>of</strong> direction <strong>of</strong><br />

principal stress. as well as ratio <strong>of</strong> shearing to dilating 0.<br />

Since we feel that leads would be more likely to be generated along<br />

lines <strong>of</strong> velocity discontinuities than along stress characteristics. it is<br />

perhaps more direct to express characteristic directions in terms <strong>of</strong> b-he<br />

stretching tensor rather than the stress state. We have<br />

148


and this relationship is given in terms <strong>of</strong> measurable quantities. From<br />

these two directions both'y and 0 may be inferred, leaving only the magnitude<br />

<strong>of</strong> the stretching undetermined.<br />

It isanticipated that the characteristic equations will also be useful<br />

for understanding flow <strong>of</strong> ice through restricted channels and other basic<br />

problems that have not yet been addressed adequately.<br />

ACKNOWLEDGMENTS<br />

We thank Max D, Coon, John F. Nye, and Roger Colony for helpful discussions<br />

on plasticity and characteristic analysis. They have helped us to<br />

avoid many <strong>of</strong> the pitfalls that we might have encountered. This work was<br />

supported by the National <strong>Science</strong> Foundation Grant OPP71-09031 to the<br />

<strong>University</strong> <strong>of</strong> <strong>Washington</strong> for the Arctic Sea Ice Study.<br />

REFERENCES<br />

Colony, R. 1975. The simulation <strong>of</strong> sea ice dynamics. In Proceedings, Third<br />

Internutima2 Conference on Port and Ocean Enginee2.ing Under Arctic Conditions,<br />

vol. 1, pp. 469-486, Institute <strong>of</strong> Marine <strong>Science</strong>, Univ. <strong>of</strong><br />

Alaska, Fairbanks.<br />

Coon, M. D., and R. S. Pritchard. 1974. Application <strong>of</strong> an elastic-plastic<br />

model <strong>of</strong> arctic pack ice. In The Coast and SheZf <strong>of</strong> the Beaufort Sea .<br />

(ed. J. C. Reed and J. E. Sater), pp. 173-193, Arctic Institute <strong>of</strong> North<br />

America, Arlington, Va,<br />

Coon, M. D., G. A. Maykut, R. S. Pritchard, D. A. Rothrock, and A. S. Thomdike.<br />

1974. Modeling the pack ice as an elastic-plastic material. <strong>AIDJEX</strong><br />

BuZZetin, 24, 1-106.<br />

Coon, M. D., R. Colony, R. S. Pritchard, and D. A. Rothrock. 1976. Calculations<br />

to test a pack ice model. In NwneKcaZ Methods in Geomechanics<br />

(ed. C. S. Desai), vol. 2, pp. 1210-1227, American Society <strong>of</strong> Civil<br />

Engineers, New York.<br />

Coon, M. D., R. T. Hall, and R. S. Pritchard. 1977. Prediction <strong>of</strong> arctic ice<br />

conditions for operations. In ProceecXngs, Ninth Offshore TechnoZogy<br />

Conference, vol. 4, pp. 307-314, Houston, Texas.<br />

149


Courant, R., and D. Hilbert. 1962. Methods <strong>of</strong>Mathematica1 Physics, vol. 2,<br />

Interscience Publishers , New York.<br />

Crandall, S. H., and N. C. Dahl. 1959. An Introduction to the Mechanics <strong>of</strong><br />

SoZids, McGraw Hill, New York.<br />

Haithomwaite, R. M. 1963. Progress in Applied Mechanics: The Prager Anniversary<br />

Volume, MacMillan, New York.<br />

Hill, R. 1950. The<br />

London.<br />

Mathematical Theory <strong>of</strong> Plasticity, Oxford Univ. Press,<br />

Hodge, P. G. 1950.<br />

Mathematics and<br />

Yield conditions in plane plastic stress.<br />

Physics, 29, 38-.<br />

JoumaZ <strong>of</strong><br />

Jenike, A. W. 1961.<br />

<strong>of</strong> Utah, 52(29).<br />

Gravity flow <strong>of</strong> bulk solids.<br />

BuZZetin <strong>of</strong> the <strong>University</strong><br />

de Jong, G. de Josselin. 1959. Statics and kinematics in the failable zone<br />

<strong>of</strong> a material. Thesis, <strong>University</strong> <strong>of</strong> Delft.<br />

McPhee, M. G. 1975. Ice-ocean momentum transfer for the <strong>AIDJEX</strong> model. <strong>AIDJEX</strong><br />

BulZetin, 29, 65-85.<br />

Marco, J. R., and R. E. Thomson. 1975. Spatially periodic lead patterns in<br />

the Canada Basin sea ice: a possible relationship to planetary waves.<br />

GeophysicaZ Research Letters, 2(10), 431-434.<br />

Marco, J. R., and R. E. Thomson. 1976. Rectilinear leads and internal motions<br />

in the ice pack <strong>of</strong> the western Arctic Ocean. JownaZ <strong>of</strong> Geophysical<br />

Research, 82 (6), 979-987.<br />

Martin, J. B. 1975. PZasticity: FmdamentaZs and General Results, MIT Press,<br />

Cambridge , Mass.<br />

Maykut, G. A., A. S. Thorndike, and N. Untersteiner. 1975. <strong>AIDJEX</strong> scientific<br />

plan. AIWEX BuZZetin, 25, 1-67.<br />

Morrison, H. L., and 0. Richmond. 1976. Application <strong>of</strong> Spencer's ideal soil<br />

model to granular materials flow. JoumzaZ <strong>of</strong> AppZied Mechanics, 98(1), 49-.<br />

Nye, J. F. 1975. Discontinuities in the <strong>AIDJEX</strong> model. <strong>AIDJEX</strong> <strong>Bulletin</strong>, 28,<br />

119-126.<br />

Parmerter, R. R., and M. D. Coon. 1972. Model <strong>of</strong> pressure ridge formation<br />

in sea ice. Journal <strong>of</strong> GeophysicaZ Research, 77(33), 6565-6575.<br />

Prager, W. , and P. G. Hodge. 1951. Theory <strong>of</strong> Perfectly Plastic SoZZ:ds, John<br />

Wiley, New York.<br />

Pritchard, R. S. 1975. An elastic-plastic constitutive law for sea ice.<br />

JoumaZ <strong>of</strong> AppZied Mechanics, 42 (ser. E, no. 2), 379-384.<br />

150


Pritchard, R. S., and R. Colony. 1974: One-dimensional difference scheme for<br />

an elastic-plastic sea ice model. In Corputational Nethods in Armlinear<br />

Nechanics (ed. J. T. Oden et al.), pp. 735-744, Uriversity <strong>of</strong> Texas,<br />

Austin.<br />

Pritchard, R. S., and R. Colony. 1976. A difference scheme for the <strong>AIDJEX</strong><br />

sea ice model. In Numerical Pethods in Geomechanics (ed. C. S. Desai),<br />

vol. 2, pp. 1194-1209, American Society <strong>of</strong> Civil Engineers, New York.<br />

Pritchard, R. S., and R. T. Schwaegler. 1975. In Proceedings, Third International<br />

Conference on Port and Oceax EndneeKng Under Arctic Conditions,<br />

pp. 513-526, <strong>University</strong> <strong>of</strong> Alaska, Fairbanks.<br />

Pritchard, R. S., M. D. Coon, and M. G. McPhee. 1976. Simulation <strong>of</strong> sea ice<br />

dynamics during <strong>AIDJEX</strong>. Journal <strong>of</strong> Pressure Vessel Technolo~y, 99<br />

(ser. J, no. 3), 491-497.<br />

Pritchard, R. S., M. D. Coon, PI. G. McPhee, and E. Leavitt. 1977. Winter<br />

ice dynamics in the nearshore Beaufort Sea. <strong>AIDJEX</strong> <strong>Bulletin</strong>, 37, 37-93.<br />

Sch<strong>of</strong>ield, A., and P. Wroth. 1968. Critical State Soil blechanics, McGraw-<br />

Hill, Ltd. , Maidenhead, England.<br />

Schwaegler, R. T., and R. S. Pritchard. 1977. <strong>AIDJEX</strong> model response to<br />

axisymmetric loadings. Paper presented at Symposium on Sea Ice Processes<br />

and Models, 6-9 Sept., <strong>University</strong> <strong>of</strong> TJashington, Seattle. Proceedings<br />

to be published by UW Press.<br />

Shield, R. T. 1955. On Coulomb's law <strong>of</strong> failure in soils. Journal <strong>of</strong><br />

Mechanics and Physics <strong>of</strong> Solids, 4, lo-.<br />

Sodhi, D. S. 1977. Ice arching and the drift <strong>of</strong> pack ice through restricted<br />

channels. Ch'REL Report 77-18, Cold Regions Research and Engineering<br />

Laboratory, Hanover, N.K., pp. 1-11.<br />

Spencer, A. J. M. 1964. A theory <strong>of</strong> the kinematics <strong>of</strong> ideal soils under<br />

plane strain conditions. Journal <strong>of</strong> Mechanics and Physics <strong>of</strong> Solids,<br />

12, 337-.<br />

151


A RESEARCH PLAN TO TEST THE <strong>AIDJEX</strong> MODEL<br />

AS AN ICE FORECASTING AID<br />

Robert S. Pritchard and Max Coon<br />

<strong>AIDJEX</strong><br />

S LJPMARY<br />

The <strong>AIDJEX</strong> ice model provides a method for describing in detail the<br />

motion and state <strong>of</strong> sea ice. The mathematical model is a momentum<br />

balance consisting <strong>of</strong> air stress, ocean stress, Coriolis force, and<br />

internal ice stress. From a calculation the velocity and deformation<br />

<strong>of</strong> sea ice may be determined, as well as the state <strong>of</strong> the ice<br />

as described by an ice thickness distribution. The input information<br />

required for the model calculation is the air stress as determined<br />

from a barometric pressure field and the boundary velocities for the<br />

domain <strong>of</strong> interest. Previous work shows that this ice model performs<br />

very well when high-quality data are available. In the present work<br />

the possibility <strong>of</strong> using this model as one component <strong>of</strong> an ice forecasting<br />

system for the Arctic is investigated. An actual ice forecast<br />

would, <strong>of</strong> course, utilize all available data, including the output<br />

<strong>of</strong> the <strong>AIDJEX</strong> model. This report examines only the quality <strong>of</strong> a<br />

calculation as part <strong>of</strong> a forecast.<br />

To use the <strong>AIDJEX</strong> model in a forecasting mode requires only that<br />

barometric pressure and boundary motions be predicted. Therefore,<br />

we have chosen to study the effects <strong>of</strong> these inputs separately.<br />

This report examines a period in late January and early February 1976<br />

for which ice model calculations have been made from a full set <strong>of</strong><br />

<strong>AIDJEX</strong> data. The calculations are used with NOAA satellite images<br />

to judge the quality <strong>of</strong> the forecast ice conditions. Maps from the<br />

National Weather Service which predict surface barometric pressure<br />

for 24-, 36-, and 48-hour intervals were used as input for the calculation,<br />

as were actual measured boundary velocities. In all simulations<br />

there was good qualitative agreement in the velocity, deformation,<br />

and stress fields and littlevariationbetween the predictions for a<br />

24- and a 48-hour period.<br />

!This report originated as an annual report, June 1077, to the Environmenta<br />

2 Research Laboratories, NCAR, Boulder, Co Zorado.<br />

153


It is notknownat present how to predict boundary motions accurately.<br />

However, a parameter study (Pritchard and Thomas, 1977) shows that<br />

errors introduced at the boundary decay in magnitude with distance<br />

from the boundary. This is true for velocity errors that have random<br />

variation in time about a zero mean. The effect <strong>of</strong> steadystate<br />

errors in the boundary motion is not known; however, it is<br />

known that they do not necessarily decay with distance.<br />

It is shown here that the <strong>AIDJEX</strong> model provides accurate predictions<br />

<strong>of</strong> ice behavior on time scales <strong>of</strong> one or two days. The<br />

pressure field predictions can be inaccurate, especially if the<br />

prediction is made for longer times; however, the errors do not<br />

dominate the behavior <strong>of</strong> the predicted ice response.<br />

It is felt that it will be necessary to deploy and operate a set <strong>of</strong><br />

data buoys to determine the drift <strong>of</strong> the ice and the barometric<br />

pressure if reliable predictions are to be made under all conditions.<br />

At present it is recommended that these buoys be deployed in a region<br />

along the entire continental shelf <strong>of</strong> the Alaskan north slope, from<br />

the shore out to approximately 800 km. These will provide accurate<br />

data from which to predict barometric pressure, as well as important<br />

information on the state <strong>of</strong> the ice at any instant so that predictions<br />

can be updated.<br />

INTRODUCTION<br />

In this work we examine the <strong>AIDJEX</strong> air, ice, and ocean models to ascertain<br />

their usefulness in forecasting the motion and condition <strong>of</strong> arctic sea ice<br />

in the nearshore portions <strong>of</strong> the Beaufort and Chukchi seas. Because we are<br />

concerned with the short-term forecasting needed to insure safe day-to-day<br />

operation <strong>of</strong> ships and marine structures, we limit the time scale to a<br />

maximum <strong>of</strong> two days, the longest period for which predictions <strong>of</strong> atmospheric<br />

conditions were available when this research was performed.<br />

In the arctic summer, the edge <strong>of</strong> the floating ice sheet that covers<br />

the Beaufort Sea usually retreats northward from the land mass, opening a<br />

coastal band <strong>of</strong> water about 200 km wide that is, for the brief time it<br />

remains open, the sea route to the Alaskan north slope. During the 1975<br />

shipping season, attempts to use this corridor for pipeline shipments<br />

were thwarted by the worst ice conditions ever recorded. For two months<br />

barges lay south <strong>of</strong> Point Wainwright waiting for the pack to move northward<br />

so that they could sail past Point Barrow and on to Prudhoe Bay.<br />

Finally, after a loss <strong>of</strong> several months and millions <strong>of</strong> dollars, most <strong>of</strong><br />

154


the barges managed to reach their destination by breaking through ice as much<br />

as a foot thick. Others turned back and their cargoes were eventually shipped<br />

overland to Prudhoe Bay. The following year, despite long-term predictions<br />

that ice conditions during late spring and early summer would be even worse<br />

than before, tugs were able to get the barges through easily.<br />

It is possible that the main reason for this surprising turn <strong>of</strong> events<br />

was the development <strong>of</strong> a new ice-breaking barge, the Arctic ChaZZenger, which<br />

allows commercial vessels for the first time to operate within rather than<br />

around the ice pack. Indeed, one might surmise that the ability to build<br />

such icebreakers will eliminate the need for ice forecasting in shipping<br />

plans. However, in fact this advance in technology increases the ne.ed for<br />

an accurate ice forecasting system. The apparent contradiction arises because<br />

no ship--not even the Soviet icebreaker Arktika, which sailed to the North<br />

Pole in December 1977--is capable <strong>of</strong> sailing on a chosen course at a chosen<br />

speed under arbitrarily heavy ice conditions. Therefore, the prediction <strong>of</strong><br />

times and locations <strong>of</strong> heavy ice conditions and high ice pressure becomes<br />

especially useful to ships operating within the ice pack.<br />

Shipping is not the only justification for improved ice forecasting techniques<br />

in the Arctic. Oil-producing wells will soon be a reality in the<br />

shallow coastal shelf <strong>of</strong>f the North Slope, an area which is exposed to active<br />

ice motion. If oil workers know in advance what the ice is going to do, they<br />

can take certain precautions (cap a well or move the platform to a safer site,<br />

for example) to prevent a disastrous oil spill or the loss <strong>of</strong> equipment and<br />

lives.<br />

On the time and space scales that we are addressing, ice motion and<br />

behavior are controlled primarily by the winds and, to a lesser extent, by<br />

the ocean currents. If these quantities can be forecast, then ice motion<br />

and state may be forecast by the same simulation technique as was developed<br />

during <strong>AIDJEX</strong> to study sea ice response (Coon et al., 1974). Although the<br />

primary region <strong>of</strong> interest to <strong>AIDJEX</strong> is the central Beaufort Sea, it has been<br />

shown (Pritchard and Schwaegler, 1975; Coon et al., 1977; Pritchard et al., 1977)<br />

that the <strong>AIDJEX</strong> model can simulate motion and state in the nearshore regions as well.<br />

The work by Coon et al. (1977) attempts to interpret the results <strong>of</strong> the<br />

simulation <strong>of</strong> ice dynamics during spring 1975 in terms useful in a forecasting<br />

15 5


situation for operations. That period corresponds to the beginning <strong>of</strong> the<br />

<strong>AIDJEX</strong> experiment, when there were few data on barometric pressure and ice<br />

movement and the ice model used in that simulation did not accurately<br />

represent the yield strength <strong>of</strong> the ice. Our current work is compared<br />

with another simulation performed for the Outer Continental Shelf Environmental<br />

Assessment Program (OSCEAP), which simulated conditions in the<br />

nearshore Beaufort Sea during the winter <strong>of</strong> 1976. During that time there<br />

were accurate data on barometric pressure and buoy motions, and the model<br />

was modified to represent accurately behavior <strong>of</strong> the ice. It has been<br />

shown that this simulation is extremely accurate in representing the motions<br />

<strong>of</strong> the ice pack throughout the region <strong>of</strong> interest (Pritchard et al., 1977).<br />

In determining how the <strong>AIDJEX</strong> model may be used as a short-term ice<br />

forecasting tool, our specific tasks have been to determine (1) how the<br />

National Weather Service weather predictions are input best as driving<br />

forces and how much error is introduced by inaccuracies in the predictions;<br />

(2) what additional motions must be specified from data buoys to provide<br />

necessary boundary conditions and other data: and (3) how to interpret the<br />

velocity, ice state, and strain and stress states to provide necessary<br />

forecasting aid.<br />

The first task is addressed in detail in this report. The second has<br />

been addressed in part by Pritchard and Thomas (1977), who determined how<br />

the boundary velocity subject to random errors with zero mean affects solutions<br />

in the interior. The third is addressed partially by describing the<br />

results <strong>of</strong> the first two, but it also has been discussed separately by<br />

Coon et al. (1977), who describe another simulation and its application to<br />

prediction <strong>of</strong> ice conditions on the time and space scales <strong>of</strong> interest.<br />

This report is taken from an annual report to the NOAA Environmental<br />

ResearchLaboratories describing initialwork to determine how effective the<br />

<strong>AIDJEX</strong> model would be at short-term ice forecasting along the north slope<br />

<strong>of</strong> Alaska (Pritchard and Coon, 1977). It is reprinted here to make the<br />

results more available to the general scientific community.<br />

156


APPROACH<br />

A baseline simulation 0-f the conditions <strong>of</strong> interest in winter, performed as<br />

part <strong>of</strong> the OCSEAP program (Pritchard et al., 19771, shows that when sufficiently<br />

accurate barometric pressure fields and boundary motions are available the<br />

ice model predicts ice motions and behavior remarkably well. Deformations<br />

and stress are also simulated with physical reality, although these have not<br />

been as thoroughly tested for accuracy as has the motion.<br />

The purpose <strong>of</strong> the present work is to learn how badly the accuracy <strong>of</strong><br />

the simulation <strong>of</strong> ice response is degraded if driving forces are not as<br />

accurate as those available for hindcasting during the <strong>AIDJEX</strong> main experiment.<br />

We have not made a thorough evaluation <strong>of</strong> the accuracy <strong>of</strong> weather prediction.<br />

Walsh (1977), using a statistical evaluation <strong>of</strong> many comparisons, concludes<br />

that weather predictions degrade seriously at 72 hours. We, on the other<br />

hand, have taken a representative case in which accurate driving forces are<br />

known and ice motions have been accurately predicted. Using this case and<br />

a set <strong>of</strong> weather predictions during this time, we make a sensitivity study<br />

which will serve as a first check on the influence on ice behavior <strong>of</strong> such<br />

errors in the weather predictions. We have chosen the period 27 January<br />

through 3 February 1976, because accurate winds and ice motions are available<br />

and because the drifting buoys and NOAA-4 satellite images show interesting<br />

ice behavior in the nearshore Beaufort Sea. In this work we show how errors<br />

in the barometric pressure field are transmitted into errors in the ice motion,<br />

deformation, and stress for the observed conditions.<br />

It is necessary not only to determine barometric pressure fields accurately<br />

so that wind stress may be found, but also to provide boundary data for accurate<br />

forecasting <strong>of</strong> ice behavior. We have integrated the model equations in time<br />

over a portion <strong>of</strong> the Beaufort Sea approximately 800 X 1200 km in size. On<br />

the landward parts <strong>of</strong> the boundary <strong>of</strong> this region we set the velocity to zero<br />

to represent the lack <strong>of</strong> motion <strong>of</strong> the land mass. Since the plasticity model<br />

admits discontinuous velocity behavior along these boundaries, this method<br />

provides a satisfactory approach even in cases in which the ice is moving very<br />

near the shore.<br />

One<br />

late the<br />

possible alternative for providing<br />

ice response throughout the Arctic<br />

boundary motion would be to simu-<br />

Basin.<br />

Then the former boundary<br />

15 7


conditions could be used everywhere. However, this approach appears to be<br />

unnecessary and rather brutish and subject to large errors in winds far from<br />

the region <strong>of</strong> interest. In the present simulations and in all previous ones<br />

we have drifting buoys located at the northern edge <strong>of</strong> our region <strong>of</strong> interest.<br />

Since we are performing hindcasts, it is satisfactory to drive the model<br />

with these observed buoy motions. However, to use the model in a forecast<br />

mode requires that either the boundary motion or the boundary traction be<br />

predicted. There appear to be several possible schemes for predicting the<br />

boundary motion. We have not yet evaluated them. The most direct that we<br />

have thought <strong>of</strong> are (1) free drift, (2) a viscous solution in an infinite<br />

space, (3) persistence, and (4) nesting the calculation inside a larger grid.<br />

Free drift has been shown to give reasonable motions much <strong>of</strong> the time<br />

during the summer (McPhee, 1977). However, Coon et al. (1977) show it to be<br />

a poor representation <strong>of</strong> ice motion at other times. Free drift has been<br />

modified by Pritchard (1977) based on the fact that if all terms in the momentum<br />

equation are averaged over a large area, the effect <strong>of</strong> the ice strength<br />

will decay with gauge length. These large-scale free-drift results have<br />

wider application than local free drift.<br />

The second technique could use solutions developed by Hibler and Tucker<br />

(1977). Although the viscous model is inadequate for describing deformation<br />

or stress and does not represent variations observed in the nearshore velocity,<br />

it is plausible that adequate far-field boundary motions could be derived by<br />

such solutions. Both <strong>of</strong> these techniques for predicting boundary motions<br />

depend on area-wide weather patterns, including the pressure field outside<br />

the region <strong>of</strong> interest.<br />

The case for persistence is well known. It is the basis for many ice<br />

forecasting schemes. A review <strong>of</strong> ice forecasting services may be found in<br />

Wittmann and Burkhart (1973, 1974).<br />

The nesting <strong>of</strong> the grid in a larger region has been used in other<br />

geophysical problems to represent far-field effects. However, such a scheme<br />

requires a sequence <strong>of</strong> simulations in which the first one determines motion<br />

in a large region with a coarse grid and subsequent simulations use these<br />

motions as boundary input with a finer grid.<br />

15 8


SENSITIVITY OF ICE RESPONSE TO DIFFERENCES IN AIR STRESS<br />

The report on the baseline simulation <strong>of</strong> ice dynamics in winter (Pritchard<br />

et al., 1977) includes a complete description <strong>of</strong> the <strong>AIDJEX</strong> model and all<br />

parameters used. The baseline results are obtained from the Run 3C calculation<br />

with a yield strength <strong>of</strong> 10' dyn cm-'. The accuracy <strong>of</strong> the barometric<br />

pressure field and corresponding geostrophic winds has been analyzed, and<br />

ocean currents evaluated.<br />

Ice motion during the period (27 January-3 February 1976) was strongly<br />

affected by the stress. During the first two days there was no motion; when<br />

it began, it was to the west. Even while the ice was motionless, the winds<br />

were moderately strong. Ice in the eastern portion <strong>of</strong> the Beaufort Sea<br />

responded later than in other areas. Vast ice regions in the nearshore were<br />

separated from the moving pack by a discontinuity. These conditions have been<br />

verified by NOAA-4 satellite images and data from drifting buoys and <strong>AIDJEX</strong><br />

camps. The satellite images were also useful in identifying and locating a<br />

flaw lead that was an important feature <strong>of</strong> the ice dynamics. Using these<br />

data in a test <strong>of</strong> the simulated response <strong>of</strong> the ice pack, it was found that<br />

the motion was represented accurately throughout the region <strong>of</strong> interest,<br />

especially in the nearshore where the velocity discontinuity occurred.<br />

In this work we have chosen to recalculate ice behavior for three days<br />

in the period: 27 January, 30 January, and 2 February. Ice motion on these<br />

days represents a full range <strong>of</strong> response seen during the whole period: on<br />

27 January, no motion even though there are winds; on 30 January, strong<br />

motion <strong>of</strong> the entire region to the west at about 20 cm per second, and strong<br />

winds; and on 2 February, the motion reversed, with the ice blown toward the<br />

east and Banks Island.<br />

We have chosen to use the predicted barometric pressure maps with verifying<br />

times at 1200 GMT for each <strong>of</strong> the three days as driving forces for the<br />

model. No other changes were made in the simulations so that we could isolate<br />

the effect <strong>of</strong> the errors in the predicted air stress. We have taken 24-,<br />

36-, and 48-hour weather predictions for the study; the 72-hour predictions<br />

we would have preferred were not available. To clarify our terminology, a<br />

24-hour prediction for 27 January means that our verifying time is 1200 GMT<br />

on 27 January and is predicted from conditions one day preceding.<br />

15 9


Figures 1-3 show the baseline air stress fields (derived from NWS<br />

analyses <strong>of</strong> surface barometric pressure which were augmented by all <strong>AIDJEX</strong><br />

data; see Appendix) compared with NWS predictions made 24, 36, and 48 hours<br />

preceding the day in each figure. We have chosen to present the driving<br />

force using the predicted air stress. It can be seen that the conditions<br />

are quite different for the three days. On 27 and 30 January the general<br />

features <strong>of</strong> the actual air stress field are represented. On the 30th it<br />

appears that the low pressure moving into the area is displaced, and so<br />

the air stress does not yet appear within the region <strong>of</strong> interest; on 27<br />

January the features observed appear even at the 48-hour prediction level,<br />

although the magnitude in the northern part <strong>of</strong> the region is not modeled<br />

correctly. On 30 January, when high winds were noted throughout the area,<br />

the predicted winds show the correct pattern everywhere, and the 48-hour<br />

prediction is actually better than either the 24- or the 36-hour prediction.<br />

This, <strong>of</strong> course, is chance and not pro<strong>of</strong> <strong>of</strong> a more reliable prediction.<br />

In general, although the main features <strong>of</strong> the air stress field are represented<br />

in these predictions, many <strong>of</strong> the details within the region are approximated<br />

poorly.<br />

To demonstrate the effect <strong>of</strong> these errors in air stress on the simulated<br />

ice response, we performed the &mulation with each <strong>of</strong> the predicted air<br />

stress fields. These calculations provide us with a direct means <strong>of</strong> seeing<br />

how the input errors are transmitted to the simulated ice response.<br />

Before turning to those results, we note that if these driving forces<br />

were observed during times when ice conditions were light enough that ice<br />

stress had no effect on the ice response (a time that occurs <strong>of</strong>ten in the<br />

summer), it would be meaningful to learn, by means <strong>of</strong> a free-drift model,<br />

how the errors were transmitted to the ice response. We have used such a<br />

model with a set <strong>of</strong> driving forces and computed the free-drift or winddriven<br />

ice velocities during the three days <strong>of</strong> interest. The results are<br />

presented in Figures 4-6. In the past we have noted (Pritchard et al., 1977)<br />

that free drift provides a poor prediction <strong>of</strong> the ice response during this<br />

eight-day period. However, this example is not aimed at evaluating<br />

the accuracy <strong>of</strong> the free-drift response, but at estimating the effect <strong>of</strong><br />

errors in the air stress during conditions when ice stress is unimportant.<br />

160


Since free drift <strong>of</strong> ice is computed as a local response, it is relatively<br />

simple to state the magnitude <strong>of</strong> the error that occurs under these conditions.<br />

There is a one-to-one relationship between the orientation <strong>of</strong> the velocity and<br />

the orientation <strong>of</strong> the air stress at a given magnitude <strong>of</strong> air stress. An<br />

error in air stress orientation <strong>of</strong> one degree is therefore transmitted into a<br />

one degree error in ice velocity orientation.<br />

The errors in ice speed related to errors in air stress magnitude are not<br />

as simple to describe, but since there is a quadratic relationship, there is.<br />

again a direct transmittal <strong>of</strong> the errors. It must be pointed out that at certain<br />

times, especially during 27 January when winds were low, the ocean<br />

currents can have an appreciable effect on free-drift ice velocities. For that<br />

time the ocean currents provide a larger input in many areas than does the air<br />

stress. Since this quantity remains fixed when driving with either observed<br />

or predicted air stress, we can expect the simulated results to be falrly<br />

consistent, and this is observed. For 27 January the simulated results using<br />

predicted air stresses, even out to times <strong>of</strong> 48 hours, are quite reasonable<br />

in most <strong>of</strong> the region. However, near Banks Island the magnitude and orientation<br />

are appreciably <strong>of</strong>f as the prediction interval increases to 48 hours.<br />

In Figure 5, where we show results for 30 January, it is seen that the<br />

general response is again reasonably accurate. However, in this case the predicted<br />

drift is not accurate between Barrow and Prudhoe Bay. Otherwise, the<br />

results show relatively good qualitative agreement.<br />

For 2 February (Fig. 6), when the winds come about and show a large<br />

variation across the region, the variation is represented quite well with the<br />

24-hour prediction and reasonably well with the 36-hour prediction, but the<br />

correlation is degraded appreciably with the 48-hour prediction. This deviation<br />

occurs because the location <strong>of</strong> the high region is wrong at the time.<br />

We have chosen to simulate the winter conditions presented by Pritchard<br />

et al. (1977) using the ice model with a squished teardrop yield surface and<br />

a yield strength <strong>of</strong> lo* dyn em-’, because it produced the best representation<br />

<strong>of</strong> ice velocity in the study performed for OCSEAP. We have modified the air<br />

stress driving forces during the selected three days to see how much effect<br />

the change has on the simulated ice response. Results are presented in<br />

Figures 7-9.<br />

161


From Figure 7 it is seen that for 27 January, when the ice was still<br />

although fairly strong winds were blowing, the effect <strong>of</strong> this error had a<br />

negligible effect on the ice response. In all cases the ice was predicted<br />

to remain at rest, and this was observed. We point out at this point that<br />

the response observed with drifting buoys agrees with the baseline values.<br />

The details <strong>of</strong> this information are given by Pritchard et al. (1977). We<br />

see in Figure 8 that the ice response predicted for 30 January tends to<br />

match the observed response. However, the width <strong>of</strong> the region that is nearly<br />

at rest along the North Slope around to Banks Island is quite different in<br />

the three cases. The predicted values from 24 and 36 hours show this region<br />

at rest to be much larger than what was observed. Although the 48-hour prediction<br />

provides a velocity field nearly identical to the observed field and<br />

baseline field, we must consider this fortuitous, and we do not expect this<br />

result to be typical. In Figure 9 weshowthe ice velocity field during 2 February,<br />

when the direction has changed. All velocity fields generated using 24-,<br />

36-, and 48-hour predicted air stresses are essentially the same.<br />

In this winter simulation when the ice strength is important, errors in<br />

the air stress do not appear to be transmitted directly into errors in the ice<br />

response. Under these conditions it appears that the predicted ice response<br />

is relatively insensitive to local errors in air stress. It is suggested that<br />

the ice responds to a spatial average <strong>of</strong> the air stress field,and so local<br />

details <strong>of</strong> the air stress are less important. A comparison with Figure 2 shows<br />

that the average air stress on 30 January was only about half the observed<br />

value, so that not only are local errors evident, but the large-scale average<br />

error over this entire region is considerable.<br />

We must be careful not to lean too heavily on the results for 30 January<br />

because <strong>of</strong> an error that occurred during the baseline calculation. As has been<br />

discussed by Pritchard et al. (1977), erroneous data were used to generate the<br />

baseline pressure map; they increased the geostrophic flows in the region <strong>of</strong><br />

the manned array to values approximately 25% greater than observed. We must,<br />

therefore, reduce the observed motions accordingly in this simulation, which<br />

would tend to reduce the size <strong>of</strong> the area brought into motion by the air stress<br />

and reduce the magnitude <strong>of</strong> the velocities within the area as well.<br />

In summary, then, we must say that, when ice stress has a significant<br />

effect on ice response, a difference in air stress between what is predicted<br />

162


and what is observed becomes less important.<br />

on free drift:<br />

currents were dominant.<br />

This is much like our comment<br />

that errors in the air stress were less important when ocean<br />

In the present case we have the same contribution<br />

from ocean currents and one other important one as well, from ice stress<br />

divergence.<br />

Now that we have evaluated the effect <strong>of</strong> air stress errors on ice velocity,<br />

we take a closer look at certain second-order quantities, such as stretching<br />

and stress.<br />

In Figures 10-12 we present the stretching tensor field during the three<br />

days we have simulated using the predicted air stress fields.<br />

In all cases<br />

there is general agreement; that is, when no deformation is occurring (as for<br />

27 January), the same condition exists in the 24-, 36-, and 48-hour predic-<br />

tions.<br />

In fact, during this day all four plots are nearly the same. Only<br />

small differences exist near the northwest boundary where the greatest<br />

stretching occurs.<br />

There the baseline stretching shows a maximum <strong>of</strong> about<br />

3% per day uniaxial opening, while the predictions all show a maximum <strong>of</strong> about<br />

2%.<br />

In Figure 11, for 30 January, we see the same general pattern <strong>of</strong> large<br />

shear in a narrow band <strong>of</strong>f the North Slope, with uniaxial opening near Banks<br />

Island.<br />

In this case, however, there is a difference between magnitudes and<br />

locations <strong>of</strong> the maxima.<br />

the development <strong>of</strong> a narrow band <strong>of</strong> shear.<br />

For example, the 24- and 36-hour predictions show<br />

Pritchard et al. (1977) have<br />

identified it as a flaw lead beginning near Barrow, running east, and curving<br />

north to Banks Island; it can be identified readily in NOAA-4 satellite images<br />

(Pritchard et al., 1977, Figs. 5-8).<br />

cell widths) north <strong>of</strong> the observed region.<br />

The feature is located about 80 km (two<br />

the eastward part <strong>of</strong> the region is at rest and not moving as in the baseline<br />

calculation.<br />

Thus, the maximum uniaxial opening occurs about 200 km from<br />

Banks Island rather than at the shore, as shown in Figure 9.<br />

We should point<br />

out, however, that the data buoy tracks (Pritchard et al., 1977) show the ice<br />

in this region to be at rest. A look at Figure lld shows that the 48-hour<br />

prediction is quite similar to the observed motions.<br />

Furthermore, as we saw in Figure 8,<br />

The baseline and all predicted stretching fields for 2 February (Fig. 12)<br />

are similar to each other. Differences are on the order <strong>of</strong> 1% per day.<br />

This<br />

16 3


condition is similar in magnitude to 27 January, when the major portion <strong>of</strong><br />

the region was not deforming.<br />

In summary, we find that the deformation field is approximated equally<br />

well by all predictions, and there is good qualitative agreement with the<br />

stretching. It is extremely important to represent the deformation field<br />

well, since it is the stretching tensor that causes mechanical redistribution<br />

<strong>of</strong> the ice state. In the present simulations we have ignored the thickness<br />

distribution for reasons explained in Pritchard et al. (1977). However, it<br />

remains important to approximate the deformations even when a perfectly plastic<br />

model is used; otherwise, we would expect significant changes in the solution<br />

when the thickness distribution is included and the yield strength is determined<br />

as a function <strong>of</strong> the thickness distribution. This statement is not true in<br />

all cases, but it is true in the present conditions, where changes in yield<br />

strength would not have a dominant effect on changing the solution.<br />

In Figures 13-15 a stress tensor field is simulated for the three days <strong>of</strong><br />

interest. All information contained in the stress tensors at each point is<br />

presented. Principal values are shown proportional to line length in the<br />

direction given, so that we are able to identify both principal values and<br />

direction in these field plots. It should be noted that the stress tensor<br />

is quite dependent on details <strong>of</strong> the deformation field and can be extremely<br />

sensitive to certain errors. By the same token, since the stress is constrained<br />

to lie within or on the yield surface, the amount <strong>of</strong> error that can be introduced<br />

to the stress state has a certain maximum. From these two conflicting<br />

ideas we are uncertain how sensitive the stress field will be to the errors<br />

in driving force.<br />

It is seen that the magnitude <strong>of</strong> the principal values and variations<br />

across the field are in general agreement in all cases, a result caused by<br />

the bound introduced by the yield surface. The comparison when the material<br />

is at rest on 27 January and 2 February is quite close. There seems to be no<br />

degradation in any case as prediction times go from 24 to 48 hours. For 30<br />

January the baseline response appears appreciably different on the western<br />

edge from all three predicted values. It is possible that our errors in the<br />

baseline air stress have again caused this deviation. We make this conjecture<br />

primarily because the three predicted fields shown in Figure 14 are so similar.<br />

In any case, in the nearshore regions <strong>of</strong> the Alaskan continental shelf the<br />

164


stresses appear to be relatively unaffected by errors in the air stress, even<br />

at 48-hour prediction times.<br />

It is not possible at the present time to ascertain how much effect on<br />

the stress field comes from the boundary motion and how much from the air<br />

stress driving force. We guess, however, that a significant portion <strong>of</strong> the<br />

driving force that affects the stress comes from the boundary conditions and<br />

that is why the stress field is relatively insensitive to errors in the<br />

applied air stress field. The number <strong>of</strong> errors caused by air stress relative<br />

to the number caused by boundary motion must, <strong>of</strong> course, be a significant<br />

function <strong>of</strong> the ice strength. During the winter, when strengths are LO8<br />

dyn cu-l, we have shown that the stress field feels the effect <strong>of</strong> the boundary<br />

motion to a great degree. During the summer, since strength is reduced to<br />

very low values and the stresses are contained within the yield surface, errors<br />

on the stress field will be due primarily to errors in air stress. At intermediate<br />

values the results lie between these two extremes.<br />

The work <strong>of</strong> Coon et al. (1977) describes how the stress is an important<br />

quantity in predicting ice conditions for operations. The stress, <strong>of</strong> course,<br />

is a large-scale stress, but we conjecture that it is directly related to<br />

forces acting between floes, and therefore to forces that would act on a ship<br />

or structure. We have been able to simulate the stresses, and they show that<br />

these stresses are relatively insensitive to errors in predicted barometric<br />

pressure fields.<br />

CONCLUSIONS<br />

The <strong>AIDJEX</strong> ice model, in previous work, was used to hindcast conditions<br />

observed during the <strong>AIDJEX</strong> main experiment, a time for which air stress fields<br />

were prescribed accurately and boundary motions were given by the observed<br />

motion <strong>of</strong> data buoys. In that test it was shown to simulate the ice dynamics<br />

accurately when the ice stress is important.<br />

In the present work we used the National Weather Service 24-, 36-, and<br />

48-hour predicted surface barometric pressure maps to determine the air stress<br />

fields. It was our desire to learn how much the solution accuracy is degraded<br />

if these predicted air stresses are used to drive the model in a forecast mode.<br />

16 5


We found no essential differences between the 24-, 36-, and 48-hour results.<br />

In all simulations performed there is good qualitative agreement in the<br />

velocity, deformation, and stress fields. Although at some times and in<br />

specific locations the velocity could be <strong>of</strong>f by nearly 10 cm sec-l (or about<br />

30%), this was not typical <strong>of</strong> the error over a large region.<br />

A statement <strong>of</strong> the magnitude as percent <strong>of</strong> error is not as illustrative<br />

as a thoughtful comparison <strong>of</strong> the entire field <strong>of</strong> results. In some locations<br />

and at some times the predicted solutions compared with observed motions even<br />

better than the baseline calculations. From these results we conclude that<br />

errors in the 24-, 36-, and 48-hour predicted barometric pressure fields<br />

introduce errors in the ice response that are no larger than errors in the<br />

model itself, and there is no difference in accuracy if 24-, 36-, or 48-hour<br />

predictions are used. It must be made clear that this conclusion has been<br />

reached primarily when ice stress is important.<br />

To determine ice response, boundary motion must also serve as input to<br />

the model. We have not yet learned how to predict these quantities, although<br />

several options have been described. In other work, however, we have determined<br />

how errors in boundary motion influence solutions in the interior. A<br />

thorough parameter study using a one-dimensional numerical solution scheme<br />

has shown that zero-mean random errors in the boundary velocity decay in<br />

magnitude with distance from the boundary (Pritchard and Thomas, 1977). This<br />

study is completed for the velocity and deformation field and has been begun<br />

for the stress field. To synthesize these results we have used a nondimensional<br />

form <strong>of</strong> the <strong>AIDJEX</strong> model that was developed in another phase <strong>of</strong> our <strong>AIDJEX</strong><br />

modeling effort. These results are also presented, since the nondimensionalization<br />

synthesizes results <strong>of</strong> the parameter study.<br />

Work has begun on learning whether or not steady-state errors in the<br />

boundary velocity can be expected to have a dominant effect on solutions in<br />

the interior, but results are not yet available. It can be said that there<br />

is no decay with distance, but it is not yet known how large these errors are.<br />

In summary, the <strong>AIDJEX</strong> model may be an accurate mathematical tool for<br />

predicting ice behavior on time scales on the order <strong>of</strong> 1-2 days. The ice<br />

model can provide accurate forecasts if barometric pressure fields, boundary<br />

motion, and ocean currents are input accurately. Accuracy <strong>of</strong> the ice<br />

166


ehavior is directly related to accuracy <strong>of</strong> these input data sets as<br />

well as choice <strong>of</strong> parameters on the ice model. It is felt that accuracy<br />

<strong>of</strong> barometric pressure predictions can and should be improved over<br />

present NWS predictions by deploying a modest array <strong>of</strong> drifting buoys.<br />

The buoys should relay pressure as well as position and be deployed in<br />

a region from shore out to approximately 800 km <strong>of</strong>f the north shore <strong>of</strong><br />

Alaska in the Beaufort and Chukchi seas during those times that accurate<br />

predictions are required. In any case, we have seen that the <strong>AIDJEX</strong><br />

model has simulated ice response accurately at a time when ice stress<br />

is important (and free-drift velocity predictions are inaccurate), and<br />

errors in the predicted barometric pressure field have introduced only<br />

modest inaccuracies in the ice response for this time.<br />

ACKNOWLEDGMENTS<br />

The work reported here is the result <strong>of</strong> contributions from many members<br />

<strong>of</strong> the <strong>AIDJEX</strong> modeling group. In the original report to the Environmental<br />

Research Laboratories we separated major contributions into appendices for<br />

proper attribution. Those whose contributions are less visible must be<br />

thanked here: Don Thomas, for making the simulations, and Lois Harris, for<br />

generating the graphics.<br />

This work dovetailed with other phases <strong>of</strong> the overall <strong>AIDJEX</strong> modeling<br />

efforts which were supported by the National <strong>Science</strong> Foundation, Division<br />

<strong>of</strong> <strong>Polar</strong> Programs, and by the Bureau <strong>of</strong> Land Management through the OCSEAP<br />

program in an interagency agreement with the National Oceanic and Atmospheric<br />

Administration. The interaction is apparent from the fact that the baseline<br />

study was funded by OCSEAP and the work determining the effect <strong>of</strong> boundary<br />

velocity errors was supported jointly by NOAA/ERL and NSF/DPP.<br />

APPENDIX<br />

SURFACE BAROMETRIC PRESSURE ERROR ANALYSIS<br />

R. A. Brown and M. Albright<br />

Since we are determining the driving force on the ice from National<br />

Weather Service prognosticated surface pressure maps, we must evaluate the<br />

16 7


error in this prediction. The error in the pressure field has several components.<br />

There is the inherent error in the accuracy <strong>of</strong> the instruments that<br />

provide the initial and verifying pressure fields. The accuracy required<br />

for NWS is no greater than 20.5 mb for individual instruments, while that<br />

for <strong>AIDJEX</strong> pressures is k0.3 mb. When the individual pressures are processed<br />

by hemispheric objective surface analysis (as is done by NWS), the analysis<br />

errors will be largest where the density <strong>of</strong> data points is low and the distribution<br />

thin. The sparseness <strong>of</strong> weather stations in the Arctic results in a<br />

lack <strong>of</strong> detail in the pressure field analysis. Since the ice responds to<br />

large-scale averages, this may affect the ice model prediction.<br />

For the persod <strong>of</strong> investigation, the average error in the NWS analysis<br />

and prognosis can be estimated from the <strong>AIDJEX</strong> surface analysis since the<br />

latter involves a more rigorous local pressure analysis and includes many<br />

stations: the buoy array, three camps, shore weather stations, and a weakly<br />

weighted border <strong>of</strong> points taken from the NWS analysis over the northern<br />

portion <strong>of</strong> the ocean. A sixth-order polynomial least squares fit to the<br />

<strong>AIDJEX</strong> data produces the surface pressure analysis and a readily obtained<br />

pressure gradient field. This contrasts to the NWS analysis, which generally<br />

contains only the shore stations in a computerized hemispheric analysis and<br />

therefore produces a much smoother field. The prognoses then use this analysis<br />

as initial values.<br />

From this comparison, the NWS maps contain errors averaging from t0.5 mb<br />

at the shore to k3.0 mb at the <strong>AIDJEX</strong> camp approximately 300 km <strong>of</strong>fshore.<br />

During the period in question, this error translates into an error in geostrophic<br />

wind (derived from the pressure gradient) <strong>of</strong> up to 50% in both the<br />

NWS surface analysis and the 24-hour prognosis. This is shown in Table 1,<br />

whic5, in a comparison betreen NWS analyses and proguoses and AIiIJZX analyses,<br />

estiulates errors for a representative point within the XIDJEX buoy array<br />

during the period covered. The 24- aiii 48-hour prognoses on the bottom line<br />

(the estimated error in the geostrophic flow derived from the different<br />

pressure fields) are taken from Walsh (1977).<br />

Surface pressure maps for 30 January 1976 are shown in Figure16 (made<br />

from NWS analyses) and Figure 17 (from <strong>AIDJEX</strong> analyses).<br />

16 8


TABLE 1<br />

COMPARISON OF ESTIMATED ERRORS, <strong>AIDJEX</strong> AND NWS<br />

24-hr 36-hr 48-hr<br />

<strong>AIDJEX</strong> NWS Prog. Prog. Prog.<br />

RMS error (mb) 0.6 3.5 3.3 4.7 6.4<br />

Mean error (mb) 0.4 2.5 2.8 3.9 5.0<br />

Vector error in G 10% 50% 50%* - 75%"<br />

(*> From Walsh (1977).<br />

REFERENCES<br />

Coon, M. D., G. A. Maykut, R. S. Pritchard, D. A. Rothrock, and A. S. Thomdike.<br />

1974. Modeling the pack ice as an elastic-plastic material. <strong>AIDJEX</strong><br />

BuZZetin, 24, 1-106.<br />

Coon, M. D., R. T. Hall, and R. S. Pritchard. 1977. Predictions <strong>of</strong> arctic ice<br />

conditions for operations. In Proceedings Ninth Offshore TechnoZogy Conference,<br />

vol. 4, pp. 307-314, Houston, Texas.<br />

Hibler, W. D. 111, and W. B. Tucker 111. 1977. Modeling pack ice as a viscous<br />

continuum: an examination <strong>of</strong> the circulation <strong>of</strong> the arctic ice cover over<br />

a two-year period. Unpublished manuscript, CRREL, Hanover, NH.<br />

McPhee, M. G. 1977. An analysis <strong>of</strong> pack ice drift in summer. Paper presented<br />

at Symposium on Sea Ice Processes and Models, 6-9 September, <strong>University</strong> <strong>of</strong><br />

<strong>Washington</strong>, Seattle. Proceedings to be published by UW Press.<br />

Pritchard, R. S. 1977. The effect <strong>of</strong> strength on simulations <strong>of</strong> sea ice<br />

dynamics. In Proceedings Fowth IntemationaZ Conference on Port and<br />

Ocean Engineering Under Arctic Conditions, pp. 494-505, Memorial <strong>University</strong><br />

<strong>of</strong> Newfoundland, St. Johns.<br />

Pritchard, R. S., and M. D. Coon. 1977. A research plan to test the <strong>AIDJEX</strong><br />

model as an ice forecasting aid. Annual Report to Environmental Research<br />

Laboratories, NOAA, contract no. 03-6-022-35330.<br />

Pritchard, R. S., and R. T. Schwaegler. 1975. Applications <strong>of</strong> the <strong>AIDJEX</strong> ice<br />

model. In proceedings Third InternationaZ Conference on Port and Ocean<br />

Engineering Under Arctic Conditions, vol. 1, pp. 513-526, <strong>University</strong> <strong>of</strong><br />

Alaska, Fairbanks.<br />

Pritchard, R. S., and D. R. Thomas, 1978. Response <strong>of</strong> sea ice to one-dimensional<br />

driving forces. <strong>AIDJEX</strong> BuZZetin, 38, 53-94.<br />

Pritchard, R. S., M. G. McPhee, M. D. Coon, and E. Leavitt. 1977. Winter ice<br />

dynamics in the nearshore Beaufort Sea. <strong>AIDJEX</strong> BuZZetin, 37, 37-93.<br />

Walsh, J. E. 1977. Ice forecasting limitations imposed by the accuracy <strong>of</strong><br />

atmospheric prediction models. <strong>AIDJEX</strong> BuZZetin, 36, 1-12.<br />

Wittmann, W. I., and M. D. Burkhart. 1973. Sea ice. Mariners Weather Log,<br />

part 1, vol. 17(3), pp. 125-134; also part 2, vol. 17(6), pp. 343-355, and<br />

part 3 (1974), vol. 18(4) , pp. 219-229.<br />

16 9


a) Base line b) 24-hour prediction<br />

c) 36-hour prediction d) 48-hour predi cti on<br />

Figure 1 . Air stress field at 12 GMT on 27 January.<br />

(a) are daily averages. Scale vector is 4 dyn cm-l.<br />

The base line vectors<br />

170


a) Base line b) 24-hour prediction<br />

c) 36-hour prediction d) 48-hour prediction<br />

Figure 2. Air stress field at 12 GMT on 30 January. The base line vectors<br />

(a) are daily averages. Scale vector is 4 dyn cm-’.<br />

171


a) Base line b) 24-hour prediction<br />

Figure 3. Air stress field at 12 GMT on 2 February. The base line vectors<br />

(a) are daily averages. Scale vector is 4 dyn cm-l.<br />

172


\ /<br />

a) Base line b) 24- hour predi ct i on<br />

c) 36-hour prediction d) 48-hour prediction<br />

Figure 4. Free-drift ice velocity field during 27 January. Scale vector is<br />

25 cm sec-l.<br />

173


g<br />

a) Base line<br />

3<br />

b) 24- hour predi cti on<br />

c) 36-hour prediction<br />

-<br />

d) 48-hour prediction<br />

Figure 5. Free-drift ice velocity field during 30 January.<br />

25 cm sec-l.<br />

Scale vector is<br />

174


a) Base line b) 24- hour predi ct i on<br />

c) 36-hour prediction d) 48-hour prediction<br />

Figure 6. Free-drift ice velocity field during 2 February. Scale vector is<br />

25 cm sec'l.<br />

175


a) Base line b) 24-hour prediction<br />

c) 36-hour prediction d) 48-hour prediction<br />

Figure 7. Ice velocity (daily displacement) field during 27 January. Scale<br />

vector is 25 cm sec-l.<br />

176


a) Base line b) 24-hour prediction<br />

d) 48-hour prediction<br />

Figure 8. Ice velocity (daily displacement) field during 30 January. Scale<br />

vector is 25 cm sec-’.<br />

177


a) Base line b) 24-hour prediction<br />

Figure 9. Ice velocity (daily displacement) field during 2 February. Scale<br />

vector is 25 cm sec'l.<br />

178


\ /<br />

a) Base line b) 24-hour prediction<br />

Figure 10. Stretching tensor (daily strain) field during 27 January. Principal values are proportional<br />

to time length in direction shown. Dashed lines indicate opening and solid lines closing. Scale vector<br />

is 8 x sec-l (approximately 8% per day).


c) 36-hour prediction<br />

d) 48-hour prediction<br />

Figure 10. (cont.) Stretching tensor (daily strain) field during 27 January. Principal values are proportional<br />

to line length in direction shown. Dashed lines indicate opening and solid lines closing. Scale<br />

vector is 8 x sec-’ {approximately 8% per day).


..<br />

a) Base line b) 24- hour prediction<br />

Figure 11. Stretching tensor (daily strain) field during 30 January. Principal values are proportional<br />

to line length in direction shown. Dashed line indicates opening and solid lines closing. Scale vector<br />

is 8 x sec'l (approximately 8% per day).


d) 48-hour prediction<br />

Figure 11. (cont. ) Stretching tensor (daily strain) field during 30 January. Principal values are proportional<br />

to line length in direction shown. Dashed lines indicate opening and solid lines closing. Scale<br />

vector is 8 x sec'l (approximately 8% per day).


J2<br />

a) Base line b) 24-hour prediction<br />

Figure 12. Stretching tensor (daily strain) field during 2 February. Principal values are proportional<br />

to line length in direction shown. Dashed lines indicate opening and solid lines closing. Scale vector<br />

is 8 x set" (approximately 8% per day).


c) 36-hour predi cti on d) 48-hour prediction<br />

Figure 12. (cont.) Stretching tensor (daily strain) field during 2 February. Principal values are proportional<br />

to line length in direction shown. Dashed lines indicate opening and solid lines closing. Scale<br />

vector is 8 x lom7 sec-I (approximately 8% per day).


) 24-hour prediction<br />

c) 36-hour prediction<br />

d) 48-hour prediction<br />

Figure 13. Stress tensor field on 27 January. Principal values (all comprehensive)<br />

are proportional to line lengths in directions shown. Scale vector<br />

is 10' dyn cm".<br />

185


L /<br />

a) Base line<br />

h) 24-hour predict ion<br />

c) 36-hour prediction<br />

I &<br />

/<br />

d) 48-hour prediction<br />

Figure 14. Stress tensor field on 30 January. Principal values (all comprehensive)<br />

are proportional to line lengths in directions shown. Scale vector<br />

is lo8 dyn cm'l.<br />

186


a) Base line<br />

b) 24-hour prediction<br />

/<br />

c) 36-hour prediction d) 48- hour predi ct i on<br />

Figure 15. Stress tensor field on 2 February. Principal values (all comprehensive)<br />

are proportional to line lengths in directions shown. Scale vector<br />

is 10' dyn cm-'.<br />

187


Figure 16. National Weather Service analysis for 1200 GMT,<br />

30 Jan. 1976. Each pressure contour is labeled with the last<br />

two digits <strong>of</strong> its value in millibars.<br />

188


8<br />

Figure 17. <strong>AIDJEX</strong> surface pressure analysis for 1200 GMT,<br />

30 Jan. 1976. Input data locations are denoted by symbols.<br />

Each pressure contour is labeled with the last two digits <strong>of</strong><br />

its value in millibars.<br />

189


FINAL DISPOSITION OF <strong>AIDJEX</strong> DATA BANK FILES<br />

Murray Stateman<br />

<strong>AIDJEX</strong><br />

The termination <strong>of</strong> the <strong>AIDJEX</strong> project also signals the end <strong>of</strong> the <strong>AIDJEX</strong><br />

data bank. All validated source data have been submitted to iational data<br />

banks for permanent retention, together with pertinent documentation. Copies<br />

will be distributed from those facilities upon the request <strong>of</strong> interested<br />

scientists. The following information may be useful to them.<br />

The data acquired during the <strong>AIDJEX</strong> main experimenthave been validated,<br />

for the most part, by the principal investigators and analysts. Copies <strong>of</strong><br />

finished data sets had been supplied to the data bank in several forms, such<br />

as unformatted binary using Fortran or Extended Fortran and in formatted SCOPE<br />

Internal Display code. Each set is in the format preferred by the analysts that<br />

created it. The units <strong>of</strong> common variables also vary with the analysts' preference--for<br />

example, time in minutes <strong>of</strong> the day, dates as year-days or <strong>AIDJEX</strong><br />

days (starting 1 January 1975), latitude and longitude in decimal degrees, and<br />

position in rectangular coordinates originating at the North Pole.<br />

Data presented to national data banks whose computer centers are not<br />

compatible with the CDC 6400 have been prepared in fixed block record size,<br />

external BCD, or even parity. Documentation is included as part <strong>of</strong> the headers<br />

and data records. Narrative descriptions <strong>of</strong> the data may be included as part<br />

<strong>of</strong> the header.<br />

Appendix 1 is a description <strong>of</strong> the data sets developed and used by the<br />

<strong>AIDJEX</strong> analysts. It had been published in preliminary form in <strong>AIDJEX</strong> <strong>Bulletin</strong><br />

No. 36 (May 1977) , pp. 203-210.<br />

Appendix 2 is a computerized list <strong>of</strong> the data transferred to the Environmental<br />

Data Service, which maintains the data banks at the National Oceanographic<br />

Data <strong>Center</strong> in <strong>Washington</strong>, D.C., and the National Climatic <strong>Center</strong> in Ashevil'le,<br />

North Carolina.<br />

190


APPENDIX 1<br />

<strong>AIDJEX</strong> DATA FILES<br />

1. Position <strong>of</strong> manned camps and buoys, in latitude and longitude vs. time<br />

Approximately 10 positions were calculated each day for each operating station<br />

using the Transit navigational satellite or the Nimbus F satellite. Data for<br />

the manned camps were taken from 10 April 1975 to 20 April 1976. Data for<br />

buoys in the Beaufort Sea were taken from 10 April 1975 to November 1977.<br />

Note that the lifetime <strong>of</strong> most buoys is about six months. These data characterize<br />

the motion <strong>of</strong> pack ice in the Beaufort Sea for all seasons <strong>of</strong> the year.<br />

A new set <strong>of</strong> 8 buoys were deployed in March 1977. Tracks <strong>of</strong> these buoys are<br />

part <strong>of</strong> the OCSEAP data. Their positions are given in three data transfers<br />

<strong>of</strong> 1977 to OCS. The last <strong>of</strong> these buoys died 7 Nov. 1977.<br />

Data are organized in a time series for each station, with a separation marker<br />

at the end <strong>of</strong> each 20-day period.<br />

Note: All these data ale in geographic coordinates--latitude and longitude.<br />

OCSEAP = Offshore Continental Shelf Environmental Assessment Program.<br />

2. Smoothed position, velocity, and acceleration for manned camps and buoys,<br />

in Cartesian coordinates<br />

Data from file 1 above have been post-processed using a Kalman filter technique.<br />

In one form--sorting on time--position, velocity, and acceleration from each<br />

operating buoy are arrayed together at three-hour intervals. In another form--<br />

sorting on station--position and velocity are given as a time series, separately<br />

for each station. A variance measure accompanies each element <strong>of</strong> data to<br />

characterize its error.<br />

3. Source data for RAMS buoys tracked by Nimbus F satellite<br />

Position data acquired from the start <strong>of</strong> Nimbus F operation in June 1975 have<br />

been provided by the NASA Goddard Space Flight <strong>Center</strong> and, after decoding<br />

and editing, have been incorporated into file 1 above. Several land-based<br />

RAMS packages are included in order to determine the temporal and spatial<br />

accuracy <strong>of</strong> the tracking system. No RAMS buoys were operational after<br />

7 Nov. 1977.<br />

4. Rotation <strong>of</strong> the manned camp floes (azimuth)<br />

The orientation <strong>of</strong> the camp floes, to which the Navigational Satellite Positioning<br />

System was aligned, was determined together with the camp position.<br />

Each camp azimuth, with respect to true north, was been smoothed for the<br />

period 10 April 1975 to 22 April 1976. Angular position and rate <strong>of</strong> rotation<br />

for all camps are given at three-hour intervals in a time-sorted data file<br />

together with error estimates for each datum. These data are also available<br />

in camp-sorted order, a separate time series for each camp.<br />

19 1


5. Ice thickness and snow depth<br />

Periodic measurements were made at various sites near the manned camps.<br />

Statistical evaluation <strong>of</strong> ice and snow conditions were made from frequent<br />

measurements areound a given site. Data are not continuous. Tabulations <strong>of</strong><br />

available data for the period 10 April 1975 to 29 June 1975 have been<br />

published in <strong>AIDJEX</strong> <strong>Bulletin</strong> No. 32 (June 1976). Subsequent data have not<br />

been organized or validated.<br />

6. Ice surface pr<strong>of</strong>ile (CRREL laser)<br />

One pr<strong>of</strong>ile <strong>of</strong> the ice surface was taken using a laser altimeter in the NASA<br />

990 as it traveled a 72 km track between two manned camps. A data point is<br />

a height above a reference plane every 0.4 m along the track. The measurements<br />

were made on 24 April 1975. There are 181,000 data points from Snow<br />

Bird (lat. 76.292ON, long. 146.082OW) to Big Bear (lat. 76.473ON, long.<br />

143.708OW).<br />

7. Landsat (ERTS) 1 and 2 images<br />

Satellite photos <strong>of</strong> the Beaufort Sea region have been obtained from the EROS<br />

Data <strong>Center</strong> for qualitative and quantitative analysis. About 1500 photos<br />

taken when visibility and cloud cover permitted are on file. Each photo<br />

covers a square region 100 miles on the side. Time periods are spring and<br />

fall 1972, 1973, and 1975.<br />

8. NOAA-4 and NOAA-5 satellite images<br />

Photos <strong>of</strong> the Arctic from Greenland to the Bering Straits were received daily<br />

from NESS since 2 Jan. 1975. Two images cover the belt between 70 and 80<br />

degrees N. latitude--that is, each photo covers a square area about 600<br />

miles on the side. Only infrared photos are available for the winter (Nov.<br />

through Jan.); both IR and visible photos are taken during the rest <strong>of</strong> the<br />

year. These are source data for examining large-scale ice movements in the<br />

Arctic as well as large-scale weather patterns. The last <strong>of</strong> the NOAA photos<br />

are dated 15 Oct. 1977.<br />

NESS = National Environmental Satellite Service, <strong>Washington</strong>, D.C. 20233.<br />

9. Coefficients for the polynomial describing surface air pressure contour<br />

From the combination <strong>of</strong> National Weather Service surface pressure maps and<br />

pressures measured at scattered points in the Beaufort Sea, two-dimensional<br />

pressure contours have been derived for every six-hour interval. These<br />

contours are a sixth-order polynomial in x and y, the grid coordinates<br />

overlying the Beaufort Sea region. The grid is rectangular and each element<br />

is 75 miles on the side. The coefficients <strong>of</strong> the polynomial are the data <strong>of</strong><br />

these file. They c.an be used to determine the pressure at any point in the<br />

area at any six-hour interval by translating latitude and longitude <strong>of</strong> the<br />

point to the grid coordinates and employing the polynomial coefficients for<br />

the time desired. The coefficients have been calculated for the period<br />

11 April 1975 to 20 April 1976.<br />

192


9A. Surface-1 eve1 ai r pressure (derived data)<br />

An alternative surface level air pressure file has been derirred fron.<br />

pressure measurements taken at the manned camps and from buoys containing<br />

pressure sensors. These data are interpolated at three-hour<br />

intervals and are combined with the geographic location <strong>of</strong> the corresponding<br />

station. Pressure, position data are given for four manned<br />

camps and for up to 14 buoys at each 3-hour interval for the period<br />

10 April 1975 to 20 April 1976.<br />

10. Geostrophic surface winds (derived)<br />

From the derived pressure data <strong>of</strong> file 9 above, geostrophic wind speed<br />

and direction have been calculated for specific points at 6-hour<br />

intervals. In the geogrid file these specific points are the grid<br />

points <strong>of</strong> a 16x16 overlay <strong>of</strong> the Beaufort Sea. The geosta file describes<br />

the geostrophic winds at the four <strong>AIDJEX</strong> manned camps and at the nine<br />

NavSat buoys.<br />

11. Pressure charts (source data)<br />

Surface and 850 mb pressure charts prepared by the National Meteorological<br />

<strong>Center</strong> for the Northern Hemisphere have been received for 0000 GMT and<br />

1200 GMT each day since April 1975. Measured pressures at the interior<br />

<strong>of</strong> the Beaufort Sea are combined with these analog data to improve the<br />

detailed accuracy <strong>of</strong> the derived pressures and winds data <strong>of</strong> files 9<br />

and 10 above. The last chart <strong>of</strong> the set is dated 14 October 1976.<br />

12. Surface-level meteorological data<br />

Meteorological instruments were in continuous operation at the <strong>AIDJEX</strong><br />

manned camps from April 1975.through April 1976. Hourly averages <strong>of</strong><br />

observed wind speed and direction at LO m and air temperatures at 2 m<br />

and 9 m above the surface have been prepared. Time series for each<br />

camp are available for the full operating period <strong>of</strong> the -in experiment.<br />

There are separation markers between each 20-day interval.<br />

13. Atmospheric inversion levels<br />

Inversion heights in the atmosphere were monitored continuously by<br />

acoustic radar at the manned camp designated as the main camp.<br />

Analog records were digitized at hourly intervals for the periods<br />

13 April-1 October 1975 and 5 November 1975-18 April 1976. As many<br />

as seven distinct inversion heights are given when they exist simultaneously.<br />

A second file has been prepared showing the average height <strong>of</strong> the dominant<br />

persistent inversion layer over a 3-hour period for every hour <strong>of</strong> the<br />

experiment .<br />

19 3


14. Ocean currents at manned camps (combined with winds and ice velocity)<br />

The manned camps served as floating platforms from which ocean currents relative<br />

to ice motion were measured continuously at depths <strong>of</strong> 2 m and 30 m.<br />

Hourly averages <strong>of</strong> ocean currents combined with hourly 10 m winds and 3-hour<br />

smoothed ice velocity (files 12 and 2) from each manned camp for the full<br />

operating period <strong>of</strong> the <strong>AIDJEX</strong> program are available on a single file. They<br />

are sorted by camp by time, with separation markers between 20-day intervals.<br />

This file is called WIIIO Wind, Ice, Ocean.<br />

15. Ocean currents measured from RAMS buoys<br />

Two RAMS spar buoys deployed <strong>of</strong>fshore in the Beaufort Sea in November 1975<br />

contained sensors which measured ocean currents at depths <strong>of</strong> 2 m and 30 m.<br />

A magnetic compass heading for the buoy and internal bearing <strong>of</strong> the sensors<br />

are given with the data at 3-hour intervals. These data have been combined<br />

with buoy positions to allow for absolute current determination. One buoy<br />

operated until 1 Oct. 1976; the other provided meaningful data only until<br />

28 March 1976.<br />

16. Oceanic mixed layer characteristics<br />

The upper ocean mixed layer is defined in depth by the point, or points, at<br />

which a rapid change in salinity occurs. This layer was measured for surface<br />

temperature, surface salinity, and depth twice daily at each manned camp.<br />

All available measurements (one per day) were published in tabular form<br />

in <strong>AIDJEX</strong> <strong>Bulletin</strong> No. 32 (June 1976). These are data for the period<br />

11 April-29 June 1975.<br />

17. Ocean depth<br />

The depth <strong>of</strong> the ocean beneath the path <strong>of</strong> the main <strong>AIDJEX</strong> camp was measured<br />

during two periods. Acoustic soundings were taken every hour from 25 May<br />

to 3 August 1975 and from 18 December 1975 to 25 April 1976. Round-trip<br />

time <strong>of</strong> sound travel is given together with interpreted depth. A file <strong>of</strong><br />

daily average depth merged with NavSat position is also available.<br />

18. Surface pressure (Val i dated) , <strong>of</strong>fshore RAMS buoys<br />

Four RAMS buoys deployed <strong>of</strong>fshore in the Beaufort Sea measured surface pressure.<br />

These measurements have been corrected for scale and sensor drift and<br />

have been smoothed and interpolated to 3-hour readings. Buoys were operational<br />

for the following periods: buoy 207, 18 March-28 August 1976; buoy 1015,<br />

23 March-30 September 1976; buoy 1245, 4 November 1975-1 October 1976; and<br />

buoy 1416, 5 November 1975-28 March 1976.<br />

In March 1977 two more RAMS buoys with pressure sensors were deployed <strong>of</strong>fshore<br />

in the OCS region. Buoy 1023 operated from 18 March to 10 July, buoy 1617<br />

from 8 March to 12 October. These pressure data were transferred to OCS.<br />

The data are sorted by buoy by time, and are merged with buoy position in<br />

latitude and longitude.<br />

194


19. Surface pressure (Val i dated) , <strong>AIDJEX</strong> camps and selected buoys<br />

NavSat systems at the four manned camps and nine NavSat buoys had pressure<br />

sensors to make detailed measurements not specifically included in the surface<br />

pressure charts <strong>of</strong> file 11 above. After appropriate corrections and calibration,<br />

these validated measurements were incorporated into the derivation <strong>of</strong><br />

area-wide geostrophic winds (file 10). These source data are available with<br />

their geographic position at 3-hour intervals. Data are sorted by station.<br />

The manned camps were operational from April 1975 to April 1976. Some <strong>of</strong><br />

the buoys (supplemented by nearby RAMS buoys) continued to operate as late<br />

as 6 December 1976. These data are also incorporated in the alternative<br />

pressure position file noted in data set 9 above.<br />

20. Weather observations , manned camps<br />

Handwritten weather notes logged daily by observers in the manned camps noted<br />

wind velocity, surface pressure, temperature, visibility, and weather. They<br />

back up the digitized data in the files noted above.<br />

21. Logbook entries, manned camps<br />

Members <strong>of</strong> the scientific groups recorded informal notes about events, equipment<br />

performance, changes or calibration <strong>of</strong> sensors, etc. Their logbooks<br />

back up the data collection procedures followed during the main experiment.<br />

22. Wind speed and direction measured by pilot balloon<br />

Pibal measurements using two tarcking theodolites were made each day at the<br />

main camp during the <strong>AIDJEX</strong> experiment. Two generations <strong>of</strong> data are stored<br />

in the data bank. Raw data consist <strong>of</strong> theodolite (angle) measurements taken<br />

at uniform intervals <strong>of</strong> time as the balloon ascended. Drag output (processed<br />

pr<strong>of</strong>iles) give zonal (to the west) and meridional (to the north) velocity<br />

components versus altitude. These data cover the period 10 April 1975 to<br />

20 April 1976 at the main camps.<br />

23. Pr<strong>of</strong>i 1 ing current meter<br />

Twice a day at each manned camp, a current meter was lowered to a depth <strong>of</strong><br />

194 m and raised at a steady rate to determine the stratification <strong>of</strong> the ocean<br />

currents. The analog outputs were digitized to show depth, speed and direction<br />

at uniform depth increments. The data bank has received 300 casts from Caribou,<br />

406 casts from Big Bear, 654 casts from Blue Fox, and 687 casts from Snow Bird--<br />

2047 casts in all to date (10 January 1978).<br />

Note that a cast may either be descending (mode 1) or ascending (mode 2). Each<br />

station usually has two casts, one <strong>of</strong> each mode. Optimally, then, there are<br />

four casts (two stations) per day per camp.<br />

195


24. Salinity and temperature versus depth at manned camps<br />

Standard STD measurements were made twice a day at each manned camp duriing<br />

the main experiment. As <strong>of</strong> 29 December 1977 the data bank received 581 casts<br />

distributed among the four manned camps. Time periods vary.<br />

25. Speci fi c humi di ty<br />

Hourly averages <strong>of</strong> vapor pressure, mixing ratio g/g, and specific humidity g/g<br />

are derived from measured dewpoint and atmospheric pressure at the main camps<br />

during the entire <strong>AIDJEX</strong> program.<br />

26. Under-ice pr<strong>of</strong>ile (subicex 1-76)<br />

Up-looking sonar was used to measure sea ice thickness in the Beaufort Sea in<br />

April 1976. A submarine track traversed the <strong>AIDJEX</strong> area centered on earibou.<br />

Tape 1 contains data obtained on a northward track from 7OoN to 75ON, a distance<br />

<strong>of</strong> 290 nautical miles. Then on a southeast track (135 degrees<br />

clockwise) to about 72.5O'M, a distance <strong>of</strong> about 195 nautical miles, data were<br />

acquired for tape 2. Tape 3 contains data from the westward track, 290 nautical<br />

miles along 72.5ON. The tapes are 7-trackY odd parity, 556bpi. They are<br />

composed <strong>of</strong> fixed length binary coded decimal records.<br />

The data include periodic geographic location, distance between sonar soundings<br />

(inversely related to speed) and surface beam diameter inversely related to<br />

depth followed by the successive distances between the under-ice surface and<br />

the calculated water surface at each sounding point along the track. Soundings<br />

are mostly about 5 feet apart along the track. A histogram is included showing<br />

number <strong>of</strong> soundings at 1-foot intervals for each more or less 10,000 feet <strong>of</strong><br />

track. Available tapes are written in three forms: EBSIDIC, SCOPE BCD, and<br />

SCOPE Internal Display code. The first is the original form, the second is<br />

SCOPE compatible, and the third is in English.<br />

27. Ocean tilt, J. R. Weber, DEMR, Canada<br />

Data were obtained at Caribou during the period 11-22 April 1976 (<strong>AIDJEX</strong> days<br />

467-488). Ocean tilt is given in microradians every 3-5 minutes.<br />

28. Atmospheric boundary layer pr<strong>of</strong>ile<br />

Data from 23 flights <strong>of</strong> an instrumented package used during February 1976<br />

at Caribou. Measurements <strong>of</strong> wind speed, wind direction, temperature, and<br />

air pressure versus height (and time) are given.<br />

29. Pr<strong>of</strong>iles <strong>of</strong> wind and temperature, spectral analysis<br />

Mean wind speed, wind direction, and temperature pr<strong>of</strong>iles collected at Big<br />

Bear in spring 1975 and at Caribou in spring 1976 from a 25 m tower, Experiment<br />

details are given in ADIJEX <strong>Bulletin</strong> No. 36 (May 1977) , pp. 157-174.<br />

NISSI data collected at 3 m and 20 m heights are included. It represents<br />

the variance <strong>of</strong> the wind speed signal in the range 0.2-1.0 Hz.<br />

196


APPENDIX 2<br />

<strong>AIDJEX</strong> DATA TRANSFERRED TO EDS<br />

The A cards give the File Identification Number, a number unique to the<br />

Environmental Data Service; station, th.e location <strong>of</strong> the data sensors; start<br />

and end dates <strong>of</strong> the data collection; and the identification <strong>of</strong> the location<br />

<strong>of</strong> the data bank copy <strong>of</strong> the data, with a coded description <strong>of</strong> the contents.<br />

The B cards give further information on the files, including the location<br />

<strong>of</strong> the operations printouts that show how the job <strong>of</strong> reformatting was done,<br />

the name <strong>of</strong> the principal analyst, the agency to which the data were sent,<br />

and when they were sent. The location <strong>of</strong> the data bank copy <strong>of</strong> the original<br />

data prior to reformatting for transfer is given, followed by the file<br />

descriptor.<br />

Comment copies <strong>of</strong> these data with documentation are obtainable at<br />

standard U.S. prices by request to the Director, NODC Data Services Division,<br />

NOAA/EDS, 2001 Wisconsin Ave. N.W., <strong>Washington</strong>, D.C. 20235; or by request to<br />

the Director, National Climatic <strong>Center</strong>, NOAA/EDS, Federal Building, Asheville,<br />

North Carolina 28801. Specify <strong>AIDJEX</strong>, and also specify <strong>AIDJEX</strong> file ID and<br />

medium onto which data are to be transcribed.<br />

Lists <strong>of</strong> A and B cards are given on the following pages.<br />

19 7


A Cards<br />

198


I -7-71327GEO<br />

A Cards (continued)<br />

773313A/P,PUh RlNG STA13750524 761206 P2320 F033A037 NCC METBUOY 1003<br />

---771311A/P~PO~ ~1hG~-STA117~O~ll__ZSiO825_ - YZ3ZO -F034AO37 NCC METalJOY 1U31<br />

77:3iZA/ P, PU5 RING STA12750303 761128 P232U FU33A037 NCC METBUOY1301<br />

-73L313AI PIPUA - RiNG-SJAl375051~-160827_____. P232Q -F036A037 NCC MEIdUOf 413.<br />

77;3;4A2 iM? A L ~ U rBd CAM? 0 750415 751006 F.2332 FOb2AO56 NCC ALIMUTH<br />

-- 732315,AZ ltlr CtLddTCA LAMP--1.-750428--760425- - P233L f uU6A05Li NCC AZIMUTH-.---<br />

77;315AZLM1Ai,uToF LAMP 2 750424 760425 P.2332 F007A058 NCC AZIMUTH<br />

- 772317AZiM~fiLudTbJ CAMPLL-75U41Uh0425- P2332- F00bA05t NCC AZIMUlH-- -<br />

77?3LJV2LOC A 1 LeSEd CAMP 0 730411 751008 P1736 FOliA210 NCC W/I/O<br />

-7?:3L9VELULITri5LA CAfiP--l--7304L16042L--P1756.F072AZlO NCC -W/11U----.-.<br />

77,323VELGiiTitSdk CAMP 2 750411 760420 P1756 F073A210 NCC W/I/D<br />

--??2321VEiuCi IiLbSd CAMP~75U4Ll.760420--PL7~b F074AZ10 NCC WILJU----- .-<br />

77;523?US1T lL\ KuMS 632 77U323 770403 1646 FO01C1603OCS 1ST 0 rRK 4<br />

__ 17:5t't?9SITIuX kAKS- 1Q35_.720304____77Q.'t03 1&46_ FOO2ClbO3OCS 1SL-Q. TdK -4<br />

77:ii5PJdiTIii kAM5 1052 770303 770403 1846 FO03C1803OC5 1ST P TRt( 4.<br />

-- 77~jL3PUbITl~~r RQMb-1064 770304-770403- --lb46 FOG4C18030C.5 1ST-a -TR& 4.<br />

77:527~~~1ri~;, RAM) 1305 770314 770403 1646 kOO5ClbO3OCS 1ST 9 THK C,<br />

--77: 3 2 3 P 11 J A 1 A L.X RAMS 16 OL- 7 7032 3--7 704 03- ___ - 1 t14 6 - F \1 Ub C 16 0 3 0 C S 1 S T--Q T R I( 4<br />

77:>23A/P Yd~lFNKAMh 1023 77031tl 770331 l84b FOO7Cl803OCS 1ST 0 TRK 4<br />

-.-7i'i33JA/k Pd~,TNkAtlb lbU7 -77dj06--770331._ __ lo46 F00dC18030CS 1ST 3 _TRK 4<br />

771322ICt l'Kut-~LSd 1U BB 720424 750424 PL)uZ> Fud2AZZO NCC LASER<br />

-77 L 3 Z 3 TEMP I tit zK 83 C A -MA LN 7 5 O.rl3-76 0418 --_P AZ> A --F 040 A 307 N C C - AC S T I C__ HY L Y.<br />

77;334TELlP lt~rcKt56 LA MklN75U413 7604 ld P1151 FuZuA307 NCC ACSTlC AVG<br />

_. 773322kA FEK rAPr(Bd CAMP--Q 7~0416- .-.-75101 ---- P2332- FOllAOSt: NCC HUMLiIITY- da<br />

77;323!jATtd VAr'ttCA CAMP 1 751103 760420 P2332 FU1LAO56 NCC HUMIDITY CA<br />

hi!*; GRID- PTL- .750411-_-750430- -LP2486 F015A314 NCC GEOlJINO -1<br />

77:3Z?CEU kA8hu 6KlD P?S 760405 7h0424 Y2486 F333A314 NCC GEUWLND<br />

--77;.3;dGEO hA*r>---l3 51.I?IN-5.-...-50311 7504-30 P24E1b -2034A314 NCC -GEOd-lNN---<br />

13 STATNS 760405 760424 P2486 F052A314 NCC GtOWIND<br />

RAMS _b3Z 770404- --170530 __ Plld4 FUllA134 OCS 2ND Q TKK 9<br />

RAMh 1035 770404 770703 Plld4 F012A134 OCS ZND 0 T2K 5<br />

.R A rib .~o5 2-7 7 040 4 - 7 1 o 703 __.Y 1164.<br />

J-OA3A134 OCS 2ND-O .TKK 3<br />

ttAr7S 1064 770404 770425 P1184 k014A134 OCS ZND U TKK 9<br />

R A fi5 J.3 05- -77d 4 0 4.-7 705 0 3--- __ I? 1184- F-015413_4 OCS ZND--U-T.RK--5<br />

RAM5 1601 770404 770703 6'1164 FOlbA134 CCS 2ND U TRK 5-<br />

tihhfi5 Ld23 -7704C)l __ -72'0703- . P1184 FOZGA134 OCS 2Ni) U T2K 5 .<br />

N2Ahih 1617 770401 770703 P1184 FOZlA134 LjCS ZND 0 T2K 5<br />

Sdt3 CAMP-0- 750411 751003 ____ Y 17 5 6.<br />

199


B Cards


B Cards (continued)<br />

201

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!