29.01.2013 Views

Proceedings Comptes rendus - Poac.com

Proceedings Comptes rendus - Poac.com

Proceedings Comptes rendus - Poac.com

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

POAC 81<br />

PROCEEDINGS<br />

COMPTES RENDUS<br />

VOL. II


To order the proceedings write to:<br />

Pour <strong>com</strong>mander les <strong>com</strong>ptes <strong>rendus</strong>, ecrire a:<br />

Prof. Bernard Michel<br />

Dep. de genie civil<br />

Universite Laval<br />

Cite universitaire<br />

Quebec, Canada<br />

G1K 7P4<br />

Reprints from this publication may be made, provided credit is given to the authors and<br />

reference is made to the <strong>Proceedings</strong> of the Sixth International Conference on Port and<br />

Ocean Engineering under Arctic Conditions, Quebec, Canada, 1981.<br />

Cette publication ne peut {ltre reproduite que si les auteurs en rec;oivent Ie credit et qu'une<br />

reference soit faite aux « <strong>Comptes</strong> <strong>rendus</strong> de la Sixieme conference internationale sur Ie<br />

genie maritime dans I' Arctique ", Quebec, Canada, 1981.


POAC 81<br />

The Sixth International Conference on Port and<br />

Ocean Engineering under Arctic Conditions<br />

Sixieme conference internationale sur Ie<br />

genie maritime dans I' Arctique<br />

Quebec, Canada<br />

July 27-31, 1981<br />

Du 27 au 31 juillet 1981<br />

<strong>Proceedings</strong><br />

<strong>Comptes</strong> <strong>rendus</strong><br />

Volume II<br />

Universite Laval, Quebec, Canada<br />

Ministere de I'Environnement, Gouvernement du Quebec


SPONSORS - PARRAINEE PAR<br />

Ministere de I'Environnement, Gouvernement du Quebec<br />

Universite Laval, Quebec, Canada<br />

CO-SPONSORS - CO-PARRAINEE PAR<br />

Department of Transport, Ottawa<br />

Ministere du Transport, Ottawa<br />

Canadian Coast Guard, Ottawa<br />

Garde cOtiere canadienne, Ottawa<br />

Department of Northern and Indian Affairs, Ottawa<br />

Ministere des Affaires indiennes et du Nord, Ottawa<br />

National Research Council of Canada, Ottawa<br />

Conseil national de recherches du Canada, Ottawa<br />

Canadian Committee on Oceanography<br />

Comite canadien sur I'oceanographie<br />

Arctic Petroleum Operators Association<br />

Association des operateurs petroliers de I'Arctique<br />

Eastern Petroleum Operators Association<br />

Association des operateurs petroliers de l'Est<br />

Canadian Society for Civil Engineering<br />

Societe canadienne de genie civil<br />

Order of Engineers of Quebec<br />

Ordre des ingenieurs du Quebec


INTERNATIONAL COMMITTEE - COMITE INTERNATIONAL<br />

Dr. P. Bruun, The Norwegian Institute of Technology, Trondheim, Norway<br />

(Chairman)<br />

Dr. T. Carstens, River and Harbor Lab., Trondheim, Norway<br />

Dr. R. Dempster, Memorial University of Newfoundland, St.John's, Nfld.<br />

Mr. E. Ernstons, Swedish Board of Maritime Works, Stockholm, Sweden<br />

Dr. K. Horikawa, University of Tokyo, Tokyo, Japan<br />

Mr. A. Juliusson, University of Iceland, Reykjavik, Iceland<br />

Dr. M. Maattanen, University of Oulu, Finland<br />

Dr. B. Michel, Universite Laval, Quebec, Canada<br />

Dr. W.M. Sackinger, University of Alaska, Fairbanks, USA<br />

Dr. P. Tryde, Denmark's Technical University, Lyngby, Denmark<br />

ORGANIZING COMMITTEE - COMITE D'ORGANISATION<br />

Dr. B. Michel, professeur de Mecanique des glaces, Universite Laval (president)<br />

Dr. Y. Ouellet. professeur de Genie maritime, Universite Laval (co-president)<br />

M. B. Harvey, sous-ministre adjoint. Environnement Quebec (tresorier)<br />

Dr. D. Carter, ingenieur-consultant, Quebec<br />

Mr. K. Charbonneau,Chief, Conference Services, National Research Council of Canada,<br />

M. J. Dery,<br />

Dr. M. Frenette,<br />

M. J.P. Godin,<br />

Mr. G.D. Hobson,<br />

Dr. O.H. Loken.<br />

Dr. R. Peters,<br />

Dr. J-L. Verrette,<br />

Ottawa<br />

Jacques Dery & Associes Inc., Montreal<br />

president. Societe canadienne de genie civil<br />

directeur regional, Garde cOtiere canadienne. Quebec<br />

Department of Energy. Mines and Resources, Ottawa<br />

Director, Environment Division, Department of Northern and Indian<br />

Affairs, Ottawa<br />

Associate Dean, Memorial University of Newfoundland, St.John's,<br />

Nfld.<br />

professeur d'Hydrodynamique, Universite Laval<br />

LADIES' COMMITTEE - COMITE FEMININ<br />

Mme Mariette Michel (presidente)<br />

Mme Ghislaine Carter<br />

Mme Monique Frenette<br />

Mme Suzanne Godin<br />

Mme Suzanne Harvey<br />

Mme Madeleine Ouellet<br />

Mme Claire Verreault<br />

Mme Marielle Verrette<br />

SECRETARIES - SECRETAIRES<br />

Mme Jeanne Roy<br />

Mme Diane Dussault


TABLE OF CONTENTS - TABLE DES MATIERES<br />

PAGE<br />

SPONSORS - PARRAINS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . IV<br />

COMMITTEES - COMITES ...... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . V<br />

PRESENTED PAPERS - ARTICLES PRESENTES<br />

OPENING SESSION - SESSION D'OUVERTURE<br />

A. Soucy, Societe d'Energie de la Environmental Impact of Hydro Plants in<br />

Baie James, Canada Northern Quebec- Vol. III<br />

B. Johansson, Canadian Marine<br />

Drilling Ltd, Canada<br />

Dome Petroleum Operations in the<br />

Beaufort Sea- Vol. III<br />

Session A1 MARINE STRUCTURES - STRUCTURES MARINES<br />

P. Bruun and G. Moe, Norwegian<br />

Institute of Technology, Norway<br />

V. Di Tella, G. Sebastiani,<br />

Tecnomare S.p.A., Italy<br />

C. Maclean, Panarctic Oils Ltd,<br />

W. Semotiuk, Engineered Urethanes<br />

Ltd., A. Strandberg and<br />

D. M. Masterson, Fenco Consultants<br />

Ltd., Canada<br />

B.R. Wasilewski, Gulf Canada<br />

Resources Inc., J. C. Bruce, Albery,<br />

Pullerits, Dickson & Associates,<br />

Canada<br />

C. A. Wortley, University of<br />

Wisconsin, U.S.A.<br />

J. V. Danys, Professional Engineer,<br />

Canada<br />

L. D. Brooks, Chevron Oil Field<br />

Research Co., U.S.A.<br />

- Preliminary titles<br />

Titres provisoires<br />

Design Criteria for Nearshore and<br />

Offshore Structures under Arctic<br />

Conditions<br />

Production System in Arctic Waters<br />

by using a Fully Integrated TSG Platform<br />

Ice Platforms with Urethane Foam Cells<br />

in the Neutral Axis Zone and their<br />

Application in Arctic Offshore Drilling<br />

Conceptual Design for a Mobile Arctic<br />

Gravity Platform<br />

Marine Piling and Boat Harbor Structure<br />

Design for Ice Conditions<br />

Offshore Structures on Weak<br />

Foundations Exposed to Large Ice Forces<br />

Ice Resistance Equation for Fixed<br />

Conical Structures<br />

39<br />

49<br />

60<br />

70<br />

80<br />

90


Session 81 NAVIGATION IN COLD REGIONS­<br />

NAVIGATION DANS LES REGIONS FROIDES<br />

J. Lewis, Arctec Incorporated, U.S.A.<br />

P. McCallister and Carl Argiroff,<br />

U.S. Army Engineer Detroit District,<br />

U.S.A.<br />

S. Skarborn, Albery, Pullerits,<br />

Dickson & Associates, Canada<br />

K. Takekuma, Mitsubishi Heavy<br />

Industries Ltd., Japan, P. Noble and<br />

A. Nawwar, Arctec Canada Ltd.,<br />

Canada<br />

G. Lijestrom and K. Lindberg,<br />

Gotaverken Arendal, Sweden<br />

D. Maxutov, U.S.S.R. Institute of the<br />

Arctic and Antarctic, and O. Kossov,<br />

Institute of Systems Studies, Moscow<br />

On the State of Commercial Arctic Marine<br />

Transportation<br />

Extension of the Navigation Season<br />

on the Great Lakes and St. Lawrence<br />

Seaway System<br />

Marine Transportation in Arctic Waters<br />

Transit Analysis for Delivery of Large<br />

Barges to Arctic Destinations<br />

Performance of Icebreaker Ymer on<br />

the Swedish Arctic Expedition "Ymer 80"<br />

Experience of Development and<br />

Introduction of Ice Free Ports for Cargo<br />

Ships under Arctic Conditions'<br />

Session C1 REMOTE SURVEILLANCE AND INSTRUMENTATION­<br />

TELEMETRIE ET TECHNOLOGIE DES MESURES<br />

J. Rossiter, Huntec (70) Ltd, Canada<br />

L. A. LeShack, LeShack Associates<br />

Ltd., U.S.A.<br />

J.D. Wheeler, Exxon Production<br />

Research Co., U.S.A.<br />

C. J. Bowley and J. C. Barnes,<br />

Environmental Research &<br />

Technology Inc., U.S.A.<br />

E.J. Langham, J.E. Glynn and<br />

D.A. Sherstone, National Hydrology<br />

Research Institute, Canada<br />

W.B. Jonasson, Petro-Canada and<br />

C. Durand, Centre de developpement<br />

des transports, Canada<br />

Remote Surveillance and Instrumentation<br />

in Sea Ice'<br />

Correlation of Under-Ice Roughness<br />

with Satellite and Airborne Thermal<br />

Infrared Data<br />

Ridge Statistics from Aerial<br />

Stereophotography'<br />

Comparison of Sea Ice Features in the<br />

Beaufort and Bering Seas Using Slar<br />

and Landsat Data<br />

Comparison of Pseudo-Parallax Effect<br />

and Cross-Correlation for the<br />

Computation of Ice Surface Velocities<br />

in Northern Waters<br />

An Ice Hazard Detection System -<br />

Preliminary Investigations'<br />

100<br />

107<br />

117<br />

136<br />

145<br />

Vol. III<br />

Vol. III<br />

156<br />

Vol. III<br />

166<br />

178<br />

Vol. III


Session A2 ICE MECHANICS - MECANIQUE DES GLACES<br />

B. Michel, Universite Laval, Canada Advances in Ice Mechanics 189<br />

T. D. Ralston, Exxon Production Plastic Limit Analysis of Ice Splitting<br />

Research Company, U.S.A. Failure 205<br />

N.K. Sinha, National Research Constant Stress Rate Deformation<br />

Council of Canada, Canada Modulus of Ice 216<br />

R.M.W. Frederking and G.W. Timco, Mid-Winter Mechanical Properties of Ice<br />

National Research Council of Canada, in the Southern Beaufort Sea<br />

Canada 225<br />

V.W. Neth and D.M. Masterson, Relationship Between In-Situ Confined<br />

Fenco Consultants ltd., Canada Compressive and Unconfined Laboratory<br />

Strength of Sea Water Ice> Vol. III<br />

Session 82 NAVIGATION IN COLD REGIONS-<br />

NAVIGATION DANS LES REGIONS FROIDES<br />

C. D. McKindra and T.C. Lutton,<br />

Coast Guard Headquarters, U.S.A.<br />

J. Sandkvist, University of Lulea,<br />

Sweden<br />

Statistical Analysis of Broken Ice<br />

Dimensions Generated During 140'<br />

WTGB Icebreaking Trails<br />

Conditions in Brash Ice Covered<br />

Channels with Repeated Passages<br />

H. Okamoto, K. Nozawa, H. Kawakami Dynamic Ice Loads and Stress Analysis<br />

and F. Yamamoto, Kawasaki Heavy on the Propeller of the Arctic Ship;<br />

Industries ltd., Japan Model Test in Ice 253<br />

T. Sasajima, Mitsubishi Heavy<br />

Industries ltd., Japan, V. Bulat and<br />

I. Glen, Arctec Canada ltd., Canada<br />

An Experimental Investigation of<br />

Two Candidate Propeller Designs for<br />

Ice Capable Vessels<br />

R. F. Carlson, J. P. Zarling and Engineering for Vessel Ice Accretion<br />

C. I. Hok, University of Alaska, U.S.A. with Particular Reference to the Alaskan<br />

Fishing Fleet 276<br />

Session C2 MARINE STRUCTURES - STRUCTURES MARINES<br />

J. l. Dery, Groupe Lavalin, Canada<br />

K.D. Vaudrey, Vaudrey & Associates<br />

Inc., and R.E. Potter, Sohio Petroleum<br />

Company, U.S.A.<br />

T. Yamaguchi, H. Yoshida,<br />

N. Yashima and M. Ando, Mitsui<br />

Engineering & Shipbuilding Co. ltd.,<br />

Japan<br />

Design of Wharves for Winter Navigation<br />

in the St. Lawrence River<br />

Ice Defence for Natural Barrier Islands<br />

During Freezeup<br />

Field Test Study of "Pack Ice Barrier"<br />

C. A. Wortley, University of Wisconsin, Dock Floats Subjected to Ice<br />

U.S.A.<br />

235<br />

244<br />

263<br />

286<br />

302<br />

313<br />

323


A. Shak, Tetra Tech Inc., U.S.A., Design Factors for Rubble Mound<br />

M.T. Czerniak and J.1. Colling Structures Under Ice and Wave Attack" Vol. III<br />

Session A3 ICE MECHANICS - MECANIQUE DES GLACES<br />

T. P. Taylor, Mobil Research and<br />

Development Corporation, U.S.A.<br />

Y. S. Wang, Exxon Production<br />

Research Company, U.S.A.<br />

N. Urabe and A. Yoshitake, Technical<br />

Research Center Nippon Kokan K.K.,<br />

Japan<br />

J. J. Kolle, Flow Research Company<br />

Kent, U.S.A.<br />

A. C. T. Chen, Exxon Production<br />

Research Company, U.S.A.<br />

P. C. Xirouchakis, Massachusetts<br />

Institute of Technology, and<br />

W. St. Lawrence, Cold Regions<br />

Research and Engineering Laboratory,<br />

U.S.A.<br />

An Experimental Investigation of the<br />

Crushing Strength of Ice<br />

Uniaxial Compression Testing of Arctic<br />

Sea Ice<br />

Fracture Toughness of Sea Ice­<br />

In-Situ Measurement and its<br />

Application -<br />

Fracture Thoughness of Ice;<br />

Crystallographic Anisotropy<br />

Transverse Pressure Effects on an<br />

Embedded Ice Pressure Sensor<br />

On the Acoustic Emission and<br />

Deformation Response of Finite<br />

Ice Plates<br />

T. S. Vinson, Oregon State University, Mechanical Properties of Low Density<br />

U.S.A., and T. Chaichanavong, Ice under Cyclic Axial Loading<br />

Kasetsart University, Thailand 395<br />

Session 83 METEOROLOGY AND OCEANOGRAPHY -<br />

METEOROLOGIE ET OCEANOGRAPHIE<br />

E.F. Roots, Environment Canada,<br />

Canada<br />

T. L. Kozo, Tetra Tech Inc., U.S.A.<br />

T.S. Murty, Institute of Ocean<br />

Sciences, M.1. EI-Sabh and<br />

J.M. Briand, Universite du Quebec<br />

a Rimouski, Canada<br />

D. O. Hodgins, Seaconsult Marine<br />

Research Ltd., and H. G. Westergard,<br />

Aquitaine Company of Canada Ltd.,<br />

Canada<br />

S. K. Liu and J. J. Leendertse,<br />

The Rand Corporation, U.S.A.<br />

Oceanography and Meteorology in<br />

the Arctic"<br />

Surface Wind Direction Anomalies Along<br />

the Alaskan Beaufort Sea Coast<br />

Influence of an Ice Layer on Storm<br />

Surge Amplitudes<br />

Internal Waves in Davis Strait and their<br />

Measurement with a Real-Time System<br />

A Three-Dimensional Model of Norton<br />

Sound Under Ice Cover<br />

332<br />

346<br />

356<br />

366<br />

375<br />

385<br />

Vol. III<br />

415<br />

423<br />

433


V. R. Neralla, ARMF, Canada and<br />

M. L. Khandekar<br />

Water Current Calculations for<br />

Modelling of Sea Ice Movement·<br />

Session C3 MARINE STRUCTURES - STRUCTURES MARINES<br />

I. Holand, The Norwegian Institute Risk Assessment of Offshore Structures<br />

of Technology, Norway Experience and Principles<br />

N. Urabe and A. Yoshitake, Technical Steel Selection System and Reliability<br />

Research Center, Nippon Kokan K.K., Analysis of Structures in Cold Regions<br />

Japan<br />

E. Eranti, State University of New York Dynamic Ice-Structure Interaction<br />

at Buffalo, U.S.A., F. D. Haynes, U.S. Analysis for Narrow Vertical Structures<br />

Army Cold Regions Research, U.S.A.,<br />

M. Maattanen, University of Oulu,<br />

Finland, and T.T. Soong, State University<br />

of New York at Buffalo, U.S.A.<br />

D.V. Reddy, Memorial University of Response of Offshore Towers to<br />

Newfoundland, P.S. Cheema, College Nonstationary Ice Forces<br />

of Trades & Technology, and<br />

M. Arockiasamy, Memorial University<br />

of Newfoundland, Canada<br />

M. Maattanen, University of Oulu, Experiences with Vibration Isolated<br />

Finland Lighthouses<br />

X. Jizu, Tianjin University, China, Dynamic Response of a Jacket Platform<br />

and B. J. Leira, Norwegian Institute Subjected to Ice Floe Loads<br />

of Technology, Norway<br />

H. Nakajima, N. Koma and M. Inoue, The Ice Force Acting on a Cylindrical Pile<br />

Technical Research Center, Japan<br />

Session A4 ICE MECHANICS - MECANIQUE DES GLACES<br />

T. Tabata, Hokkaido Universiy, and<br />

K. Tusima, Toyama University, Japan<br />

H. Saeki and A. Ozaki, Hokkaido<br />

University, and Y. Kubo,<br />

CR Engineering Laboratory, Japan<br />

J.-P. Nadreau and B. Michel,<br />

Universite Laval, Canada<br />

K. Cederwall, University of Lulea,<br />

Sweden<br />

P. R. Johnson, P.E., Consulting<br />

Engineer, U.S.A.<br />

Friction Measurements of Sea Ice on<br />

some Plastics and Coatings<br />

Experimental Study on Flexural<br />

Strength and Elastic Modulus of Sea Ice<br />

Creep of S2 Ice Beams and Plates<br />

Behaviour of a Reinforced Ice-Cover<br />

with Regard to Creep<br />

The Reaction of a Floating Ice Sheet<br />

to Simple Loads and Certain Classes<br />

of Vehicles and Machines<br />

Vol. III<br />

444<br />

462<br />

472<br />

480<br />

491<br />

502<br />

517<br />

526<br />

536<br />

548<br />

562<br />

571


IN VOLUME II<br />

DANS LE VOLUME II<br />

Session 84 SEA ICE CONDITIONS - CONDITIONS DE LA GLACE DE MER<br />

E. Leavitt, Intera Environmental<br />

Consultants Ltd., Canada, J. Sykes,<br />

University of Waterloo, and T.T. Wong,<br />

Intera Environmental Consultants Ltd.,<br />

Canada<br />

A Sea Ice Model Developed For Use<br />

in a Real Time Forecast System<br />

Pages<br />

R.T. Lowry, J.T. Sutton, G. J. Wessels A Sea Ice Model Developed for Use<br />

and W.C. Jefferies, Intera Environ- Real Time Forecast System, Part II:<br />

mental Consultants Ltd., Canada Extraction of Imaging Radar Data 589<br />

D. J. Agerton, Shell Oil Company,<br />

U.S.A.<br />

R. S. Pritchard and M. D. Coon,<br />

Flow Research Company, U.S.A.<br />

R. Colony and A. S. Thorndike,<br />

University of Washington, U.S.A.<br />

Large Winter Ice Movements in the<br />

Nearshore Alaskan Beaufort Sea<br />

Canadian Beaufort Sea Ice<br />

Characterization<br />

Sea Ice Strains During 1979<br />

Session C4 MARINE STRUCTURES - STRUCTURES MARINES<br />

M. Metge, B. Danielewicz and<br />

R. Hoare, Dome Petroleum Ltd.,<br />

Canada<br />

J. D. Wheeler, Exxon Production<br />

Research Company, U.S.A.<br />

On Measuring Large Scale Ice Forces;<br />

Hans Island 1980<br />

Probability Distributions for Structure<br />

Loading by Multiyear Ice Floes<br />

A. B. Cammaert, Acres-Santa Fe Inc., Impact of Large Ice Floes and Icebergs<br />

and G. P. Tsinker, Acres Consulting on Marine Structures<br />

Services Ltd., Canada 653<br />

M. Rojansky and B. C. Gerwick, Failure Modes and Forces of Pressure<br />

University of California, U.S.A. Ridges Acting on Cylindrical Towers 663<br />

A. Prodanovic, Exxon Production Upper Bounds of Ridge Crushing<br />

Research Co., U.S.A.<br />

pressure on Structures' Vol. III<br />

Session AS MARINE FOUNDATIONS AND SCOUR -<br />

FONDATIONS MARINES ET AFFOUILLEMENTS<br />

H. Kivisild, Fenco, Canada<br />

Marine Foundations'<br />

Vol. III<br />

G. R. Pilkington, Dome Petroleum Methods of Determining Pipeline<br />

Limited, and R. W. Marcellus, Canada Trench Depths in the Canadian<br />

Marine Engineering Ltd., Canada Beaufort Sea<br />

674<br />

581<br />

599<br />

609<br />

619<br />

629<br />

643


R. Abdelnour and D. Lapp, Arctec<br />

Canada Ltd., S. Haider and<br />

S.B. Shinde, Esso Resources Canada<br />

Ltd., and B. Wright, Gulf Canada,<br />

Canada<br />

R. Lien, Continental Shelf Institute,<br />

Norway<br />

Model Tests of Sea Bottom Scouring<br />

Sea-Bed Features in the Blaaenga<br />

Area, Weddell Sea, Antarctica<br />

T.R. Chari, Memorial University of<br />

Newfoundland, and S.M. Abdel-Gawad,<br />

Static Penetration Resistance of Soils<br />

University of Windsor, Canada 717<br />

H. Youssef, University of Montreal,<br />

and R. Kuhlemeyer, University of<br />

Calgary, Canada<br />

Dynamic and Static Creep Testing<br />

of Ice and Frozen Soils<br />

Session 85 SEA ICE CONDITIONS - CONDITIONS DE LA GLACE DE MER<br />

W. M. Sackinger, University of Alaska, A Review of Technology for Alaskan<br />

U.S.A. Offshore Petroleum Recovery 735<br />

R.G. Sisodiya, Gulf Research and<br />

Development Co., and K.D. Vaudrey,<br />

Vaudrey & Associates, U.S.A.<br />

Beaufort Sea First-Year Ice<br />

Features Survey - 1979<br />

D.F. Dickins, DF Dickins Engineering, Multi-Year Pressure Ridge Study<br />

and V.F. Wetzel, Suncor Inc., Canada Queen Elizabeth Islands<br />

L. Wolfson and W. M. Evans, ARCO<br />

Oil and Gas Company, U.S.A.<br />

J. R. Kreider and M. E. Thro,<br />

Shell Development Company, U.S.A.<br />

G. F.N. Cox, US Army CRREL, and<br />

W.S. Dihn, Sea Ice Consultants,<br />

U.S.A.<br />

Session C5 WAVE AND ICE MECHANICS<br />

HOULE ET MECANIQUE DES GLACES<br />

Ice Studies Aid in the Successful<br />

Completion of the Norton Sound<br />

C.O.S.T. Well<br />

Statistical Techniques for the Analysis of<br />

Sea Ice Pressure Ridge Distributions<br />

Summer Ice Conditions in the Prudhoe<br />

Bay Area, 1953-75<br />

J. Ploeg, National Research Council On the Importance of Defining Wave<br />

of Canada, Canada Climates 809<br />

P. F. Andersen, Consulting Engineer, Surface Agitation in Ice Prone Waters<br />

Canada 820<br />

A. Lachapelle, Atmospheric Winds and Waves Lancaster Sound<br />

Environment Service, Canada 830<br />

688<br />

706<br />

726<br />

755<br />

765<br />

776<br />

789<br />

799


D. Carter, Consultant, Y. Ouellet,<br />

Universite Laval, and P. Pay,<br />

Transport Canada, Canada<br />

B.D. Pratte and G.W. Timco, National<br />

Research Council of Canada, Canada<br />

Fracture of a Solid Ice Cover by<br />

Wind-Induced or Ship-Generated<br />

Waves<br />

A New Model Basin for the Testing of<br />

Ice-Structure Interactions<br />

Session A6 SEA ICE DRIFT - DERIVE DES GLACES DE MER<br />

W.D. Hibler III, U.S. Army Cold<br />

Regions Research, U.S.A., I. Udin<br />

and A. Ullerstig<br />

W.B. Tucker III and W.D. Hibler III,<br />

U.S. Army Cold Regions Research,<br />

U.S.A.<br />

R. Zorn, Danish Hydraulic Institute,<br />

and H. H. Valeur, Danish<br />

Meteorological Institute, Denmark<br />

R. R. Rumer, A. Wake and<br />

S-H. Chieh, State University of New<br />

York at Buffalo, and R. D. Crissman,<br />

GAl Consultants Inc., U.S.A.<br />

W.W. Denner, Memorial University of<br />

Newfoundland, Canada<br />

Modeling Mesoscale Ice Dynamics<br />

Using a Viscous Plastic<br />

Constitutive Law·<br />

Preliminary Results of Ice Modeling<br />

in the East Greenland Area<br />

Pack Ice Drift and Weather Impact<br />

Development of an Ice Transport<br />

Model for Great Lakes Application<br />

Numerical Modeling of Labrador Pack<br />

Ice Dynamics·<br />

Session 86 OIL SPILLS - POLLUTION PAR LE PETROLE<br />

E. Palosuo, University of Helsinki,<br />

Finland<br />

A. Kovacs, U.S.A. CRREL,<br />

R. M. Morey, Morey Research Co. Inc.,<br />

D. F. Cundy, U.S. Coast Guard Res.,<br />

and G. Dicoff, U.S.A. CRREL, U.S.A.<br />

J. D. Malcolm, Memorial University of<br />

Newfoundland, and A. B. Cammaert,<br />

Acres-Santa Fe Incorporated, Canada<br />

G. A. Robilliard and M. Busdosh,<br />

Woodward-Clyde Consultants, U.S.A.<br />

K. Horikawa and N. Mimura,<br />

University of Tokyo, Japan<br />

The Biologically Important Areas<br />

in the Arctic Ocean<br />

Pooling of Oil under Sea Ice<br />

Movement of Oil and Gaz Spills under<br />

Sea Ice<br />

Need for Real World Assessment of the<br />

Environmental Effects of Oil Spills in<br />

Ice-Infested Marine Environments<br />

Environmental Aspects of Heated Water<br />

Discharged from Coastal Power Stations<br />

843<br />

857<br />

Vol. III<br />

867<br />

879<br />

892<br />

Vol. III<br />

902<br />

912<br />

923<br />

937<br />

945


Session C6 INTERACTION BETWEEN ICE AND SHORE­<br />

INTERACTION ENTRE LA GLACE ET LES COTES<br />

J-C. Dionne, Universite Laval, Canada<br />

J. R. Harper and EH Owens,<br />

Woodward-Clyde Consultants,<br />

Canada<br />

L'action des glaces sur les littoraux<br />

Analysis of Ice-Override Potential Along<br />

the Beaufort Seacoast of Alaska<br />

A. Kovacs and D. S. Sodhi, Sea Ice Piling at Fairway Rock,<br />

U.S. CRREL, U.S.A. Bering Strait, Alaska: Observations<br />

and Theoretical Analysis 985<br />

A. Kovacs and G.F.N. Cox, Norton Sound Grounded Rubble Fields<br />

U.S. CRREL, U.S.A. and Shore Ice Pile-Ups* Vol. III<br />

B.W. Graham, Esso Resources Ice Rubble Field Stability*<br />

Canada Ltd., Canada, and S.B. Shinde Vol. III<br />

Session A7 ICEBERGS -<br />

P. Ball, H. A. Gaskill, Memorial<br />

University of Newfoundland, and<br />

R.J. Lopez, Canada<br />

D. V. Reddy and P. S. Cheema,<br />

Memorial University of Newfoundland,<br />

Canada<br />

S. D. Smith, Bedford Institute of<br />

Oceanography, and E. G. Banke,<br />

Martec Ltd, Canada<br />

R. T. Lowry, Intera Environmental<br />

Consultants Ltd, Canada, and<br />

J. S. Miller<br />

Environmental Data Requirements for<br />

a Real Time Iceberg Motion Model*<br />

Simulation of Shapes of Icebergs<br />

and their Impact Probabilities *<br />

A Numerical Model of Iceberg Drift<br />

Iceberg Mapping in Lancaster Sound<br />

with Synthetic Aperture Radar*<br />

T. R. Chari and H. P. Green, Memorial Iceberg Scour Studies in Medium<br />

University of Newfoundland, Canada Dense Sands 1012<br />

Session B7 ICE CONDITIONS - CONDITIONS DE LA GLACE<br />

J. D. Miller, Petro-Canada, Canada<br />

M. Lepparanta and E. Palosuo,<br />

University of Helsinki, Finland<br />

R. W. Reimer, J. C. Schedvin and<br />

R. S. Pritchard, Flow Research<br />

Company, U.S.A.<br />

A Sensitivity Analysis of a Simple<br />

Model of Seasonal Sea Ice Growth<br />

Studies of Sea Ice Ridging with<br />

a Ship-Borne Laser Profilometer<br />

Chukchi Sea Ice Motion<br />

955<br />

974<br />

Vol. III<br />

Vol. III<br />

1001<br />

Vol. III<br />

1020<br />

1031<br />

1038


P. McComber, Universite du Quebec<br />

a Chicoutimi, Canada<br />

J.-L. Laforte, P. C. Luan and J. Druez,<br />

Universite du Quebec a Chicoutimi,<br />

Canada<br />

W. R. McLeod, Marathon Oil<br />

Company, U.S.A.<br />

Numerical Simulation of Ice Accretion<br />

Using the Element Method<br />

The Effects of an Electric Field on the<br />

Microstructure and Mechanical<br />

Properties of Glaze and Rime<br />

Atmospheric Superstructure Ice<br />

Accumulation Measurements<br />

Session C7 ICE CONTROL MEASURES -<br />

METHODES DE CONTROLE DES GLACES<br />

J.W. Smith, University of Toronto,<br />

a.M. Kaustinen and R. O'Caliaghan,<br />

Polar Gas Project, and F. Brennan,<br />

Canada<br />

I. K. Hill and A. B. Cammaert, Acres<br />

Consulting Services, and D. R. Miller,<br />

Arctic Pilot Project, Canada<br />

K. Haggkvist, University of Lulea,<br />

Sweden<br />

Arctic Marine Heat Transfer Experiment<br />

for the Polar Gas Project><br />

A Laboratory Study of Heat Transfer<br />

to an Ice Cover from a Warm Water<br />

Discharge<br />

Combination of a Sinking Warm Water<br />

Discharge and Air Bubble Curtains for<br />

Ice Reducing Purposes<br />

1047<br />

1057<br />

1067<br />

Vol. III<br />

1094<br />

1104<br />

G.D. Fonstad, Alberta Environment, The Explosive Demolition of Floating<br />

R. Gerard and B. Stimpson, University Ice Sheets<br />

of Alberta, Canada 1114<br />

D. B. Coveney, National Research<br />

Council of Canada, Canada<br />

Cutting Ice with "High" Pressure<br />

Water Jets 1124


E. Leavitt<br />

J. Sykes*<br />

T.T. Wong<br />

A Sea Ice Model Developed For Use In A<br />

Real Time Forecast System<br />

INTERA Environmental<br />

Consultants Ltd.<br />

Calgary, Alberta<br />

Canada<br />

ABSTRACT: An ice mechanics model developed by INTERA Environmental Consultants<br />

for unconsol idated pack ice is described. This model is intended for use in a<br />

real time forecast system in support of winter dril I ing operations in the Beaufort<br />

Sea.<br />

The model solves the momentum balance for sea ice using the Galerkin<br />

finite element method. The momentum equation follows an Eulerian formulation and<br />

Includes internal ice stress, air and water stresses, Coriol is force, inertial<br />

force and ocean ti It terms. The rodel uses a plastic constitutive law with a<br />

normal flow rule and viscous closure at small strain rates. Boundary conditions<br />

can be specified using velocity or strain rate <strong>com</strong>ponents.<br />

The ice thickness distribution is described using a four category rodel<br />

wh i ch a I lows red i str i but i on of ice between categor i es due to both thermodynam i ca I<br />

and mechanical effects. Ice strength is calculated as a function of the ice<br />

th I ckness d i str i but i on and the red i str i but I on funct I on. The th i ckness<br />

distribution equation is solved using a Lagrangian formulation which enables the<br />

tracking of ice features and avoids the requirement of specifying a numerical<br />

dispersion term.<br />

A sample rode I run demonstrates how the finite element discretization<br />

can be matched to irregular boundary shapes. The use of the Lagrangian tracking<br />

scheme is illustrated by <strong>com</strong>paring the calculated positions of a data buoy with<br />

its observed trajectory.<br />

*Dr. Sykes is an assistant professor in the Department of Civil Engineering,<br />

University of Waterloo, Waterloo, Ontario<br />

581


1. I NTRODUCTI ON<br />

Petroleum industry dri II ing operations outside the fast-ice zone in the<br />

Beaufort Sea have been restricted to the ice-free season. However, Dome Petroleum<br />

is evaluating dri Iiship designs which would be capable of operating in winter ice<br />

conditions. I n a cooperat ive program, Dome and the Government of Canada have<br />

funded the development of an ice mechanics model for use in a forecast system to<br />

support year round dril ling operations.<br />

The objective of the joint program was to develop a model that would<br />

provide accurate 24 hour forecasts of ice velocities, ice convergence, and ice<br />

deformation at specific sites with high resolution. The 24 hour period is<br />

specified in order to provide sufficient warning for the drillship to be<br />

disconnected from the hole in an orderly fashion. An additional requirement was<br />

to track the location of hazardous features such as multiyear floes or areas of<br />

heavy ridging.<br />

The model ing effort was spl it into two phases: INTERA was given<br />

responsibi I ity for developing a fine scale site specific model and the Atmospheric<br />

Environmental Service (AES) was to develop a regional scale ice model that would<br />

be used to predict boundary conditions for the site specific model. The exact<br />

domain of each model was to be determined during model development. An important<br />

constraint on both models was that they be capable of producing forecasts using<br />

readily available input data.<br />

Data for model val idation were =llected during December 1979 in the<br />

Beaufort Sea. The measurement program included deployment of 5 buoys by Dome<br />

which monitored position and atmospheric pressure, and ice observations using the<br />

AES and Canadian Center for Remote Sensing airborne radars. Analysis of the radar<br />

data is described in a <strong>com</strong>panion paper [11. Ocean currents were also measured at<br />

one buoy and the Frozen Sea Research Group =nducted a CTD survey of the region in<br />

I ate November.<br />

The mode ling proj ect was <strong>com</strong>p I eted in March 1981. The formu I at i on of<br />

the site specific model and the solution technique adopted are described in this<br />

paper. A sample ice model run using the data is included to illustrate the<br />

capabil ities of the model.<br />

2. MODEL FORMULATION<br />

The modeling of the mechanical and thermodynamic behaviour of sea ice as<br />

a =ntinuum requires equations relating the forces acting upon the ice as well as<br />

equations describing the resultant redistribution of the ice (such as ridging and<br />

rafting). Fol lowing [21 the momentum equation used to describe the ice motion is:<br />

582


Boundary velocities were set equal to zero along the coast and zero<br />

norma I stra in rate boundar i es were assumed on the ice- ice boundary (F i gure 1).<br />

Twenty time steps that increased in length from 900 seconds to 3 hours were used<br />

in this simulation.<br />

The geostrophic winds were interpolated for the grid positions using a<br />

surface pressure distribution prepared by AES using avai lable buoy and shore<br />

stat ion pressure data <strong>com</strong>b i ned wi th the Canad ian Meteorolog ica I Center surface<br />

ana I ys is. Ocean currents at the bottom of the mixed I ayer were assumed to be<br />

zero.<br />

Values of drag coefficient and turning angles were 0.0008, 0.005, 25°<br />

and 20° respectively for the air and water stress calculations. The Initial ice<br />

th i ckness proport ions were assumed constant over the mode I doma i n and were set<br />

equal to 0.05, 0.05, 0.60 and 0.30 for the open, thin, flat and rubble categories,<br />

in that order. The th i ckness I eve I s for the four categor i es were 0.0 to O. 1 m,<br />

0.1 m to 0.5 m, 0.5 to 0.9 m and 0.9 to 20 m. These thicknesses are based on a<br />

I imited set of observations and are only approximate.<br />

As indicated in Figure 1, one of the thickness nodes was positioned at<br />

the location of a Dome buoy. The predicted trajectory and velocity <strong>com</strong>ponents are<br />

<strong>com</strong>pared with the observed va lues in Figures 2 to 4. I n genera I, af ter the<br />

initial six hours of simulation, excellect agreement is achieved in the y­<br />

<strong>com</strong>ponent of the velocity. However, the model tends to over-predict the x­<br />

<strong>com</strong>ponent.<br />

There are severa I poss i b I e reasons for th i s discrepancy. First I y, the<br />

reduction of the assumed drag coefficients and/or an increase in the turning angle<br />

for air stress would reduce the xvelocity <strong>com</strong>ponent and provide a better match<br />

with the observations. Secondly, examination of the wind data showed that the x<br />

<strong>com</strong>ponent of the air stress was large across the entire model domain. The large<br />

fetch would require a very high ice strength to balance this air stress if the ice<br />

is to remain relatively motionless. If the ice was less <strong>com</strong>pact than assumed in<br />

this simulation in the western half of the modeled area the effective fetch would<br />

be reduced. Consequently, sma I ler ice velocities would be calculated.<br />

5. SUMMARY<br />

This paper demonstrated that the INTERA fine scale ice model is capable<br />

of predicting ice velocities, areas of convergence and ice deformation. The model<br />

can also forecast trajectories of specific ice features. Initial tests indicate<br />

that model predictions reflect the observed behaviour. Further testing is being<br />

continued to cal ibrate the model parameterizations with the data from December<br />

1979.<br />

586


Besides providing forecast support for dril I ing operations the forecasts<br />

could also provide useful input to transportation systems. For transportation the<br />

emphasis would be to forecast areas of convergence and ice deformation for input<br />

into route selection.<br />

6. ACKNOWLEDGEMENT<br />

Th is project was supported under a contract from the Department of<br />

Supp lies and Serv ices, Government of Canada. Add i tiona I support was prov i ded by<br />

Department of the Environment, Dome Petroleum Limited, Ministry of Transport and<br />

Energy Mines and Resources.<br />

The authors acknowledge the efforts of the many INTERA co-workers who<br />

have worked together on this project to ensure its success.<br />

REFERENCES<br />

1. Lowry, R.T., J.T. Sutton and G.J. Wessels. 1981. A Sea Ice Model Developed<br />

for Use in a Real Time Forecast System, Part I I: Extraction of Imaging<br />

Radar Data. To be presented at The Sixth Conference on Port and Ocean<br />

Engineering under Arctic Conditions, Quebec City.<br />

2. Rothbrock, D.A. 1975. The Steady Drift of an In<strong>com</strong>pressible Ice Cover in the<br />

Arctic Ocean. CI imate of the Arctic.<br />

3. Coon, M.D. 1980. A Review of AIDJEX Model ing. Sea Ice Processes and Models,<br />

Ed. R.S. Pritchard. University of Washington Press. 12-27.<br />

4. Hibler III, W.D. 1979. A Dynamic Thermodynamic Sea Ice Model. Journal of<br />

Physical Oceanography. Vol. 9: 815-846.<br />

5. Reimer, R., Pritchard, R., and M. Coon. 1980. Consistent Reduction of Ice<br />

Thickness Distribution to a Few Categories. Prepared for INTERA<br />

Environmental Consultants Ltd. by FLOW Research Company. 29 pp.<br />

6. Pritchard, R.S. and M.D. Coon. 1981. Four Component Ice Characterization for<br />

the Southern Canad i an Beau fort Sea. To be presented at the Sixth<br />

Conference on Port and Ocean Engineering under Arctic Conditions, Quebec<br />

City.<br />

7.<br />

588<br />

Irons, B.M., and R.C. Tuck.<br />

Computer Iteration.<br />

Engineering. Vol. 1:<br />

1969. AVers i on of the Aitken Acce I erator for<br />

Internat iona I Journal for Numer ical Methods in<br />

275-278.


1. INI'ROJXJCTIOO<br />

An ice mechanics !1Ddel has teen developed by Intera Environmental<br />

Consultants Ltd. as part of a co-operative program between the Atr.'Dshperic<br />

Environment Service (AES) and rolE Canada (Leavitt et al., 1981). 'lliis I!Ddel is<br />

intended to te used in a real time forecast system, in support of drilling<br />

operations in winter pack ice in the Beaufort Sea. 'llie data needed to<br />

initialize, operate and update the !1Ddel can te divided into two categories.<br />

First, there are atmospheric and oceanographic data, such as winds,<br />

ter.peratures, currents, etc. Second are the ice data, sud! as type,<br />

concentration, motion, ridging, etc. 'lliis paper deals with the use of imaging<br />

radars to supply these ice data, necessary for the !1Ddel.<br />

Imaging radars were d!osen for this program tecause they provide ice<br />

data at an appropriate resolution, independent of light and atmospheric<br />

conditions. Further, X-band radars yield !!Ore information on ice type than do<br />

optical sensors or longer wavelength radars, especially if the resolution (data<br />

rate) is COIlParable (LcMry et al., 1979). 'llie techniques were developed for the<br />

analysis of radar imagery of ice to provide data for a !1Ddel. 'lliis paper is an<br />

attenpt to descrite the techniques evolved, and outlines the developments needed<br />

before radar data can be used in operational !!Odelling work.<br />

2. Dl\.TA SETS<br />

2.1 SLlIR Dl\.TA<br />

'llie SLlIR data used in this analysis were collected using the AES<br />

AN/APS 94E SLlIR. 'Ihe SLlIR data, with mud! lCMer resolution but mud! larger<br />

swath width (100 krn per side, both sides operating), were intended to te used<br />

with a "Regional Scale" !1Ddel. Four flights were COIlPleted, to cover the same<br />

two periods for whid! SAR data were oollected, naMely, 2, 6, 14, and 16 Decel!ber<br />

1979. Figure 1 shOlolS the approximate extent of the area imaged by the SLlIR.<br />

Also shown is the polar stereographic grid that was used in this I!Ddelling<br />

effort (See Leavitt et al., 1981).<br />

2 • 2 SAA DA.TA<br />

SAR data were oollected with the SAR-S80 system (Inkster et al., 1979)<br />

operated in the X-band (0.032 m wavelength), wide swath !!Ode. In this !!Ode, 22<br />

km of slant range data are collected on each pass of the aircraft. 'llie<br />

resolutioo cell of this radar is approximately square, and has an area of just<br />

over 3 m 2 • Data were oollected so that a ground range of just CNer 24 kIn was<br />

covered.<br />

'lWo !!Odelling periods were established for the SAR data: 2 to 6<br />

Decel!ber 1979, and 14 to 17 December 1979. An area of approximately<br />

10 000 km 2 was imaged 00 each day of the two periods, using four flight lines<br />

of just over 100 kIn in length. A nominal 20% overlap was planned, to allew for<br />

drift in the navigatioo. The areas covered are shown in Figure 2. Most passes<br />

had one end extended well onto land to aid in geometric calibratioo of the<br />

image. Because of poor image quality, data fran DecentJer 2nd were discarded.<br />

590


A second program was developed to find the latitude and longitude of<br />

al¥ point on the inage. '!his nade use of measurements, fran the inagery, of the<br />

distance to the biO closest timing narks. By o.:xrparing the position of points<br />

on tw:> different days, ice IlDtion vectors could t:e calculated. Further, by<br />

o.:xrparing the calculated position of the same point on biO different passes, INS<br />

errors could t:e observed. '!his was ir.portant t:ecause self-consistent radar<br />

r.nsaics could not t:e constructed for all data sets. By cbserving the "notion"<br />

of stationary ice features, including land, confidence limits were estimated for<br />

IlDtion vectors.<br />

3.2 SAR GIDlETRIC CALIBRATIrn<br />

Before SAR inagery can t:e calibrated geanetrically, using the CXlI!p.lter<br />

program, the following data nust t:e known for eam pass:<br />

1. Starting latitude and longtitude;<br />

2. Final latitude and longtitude;<br />

3. Flying height;<br />

4. Time delay and record interval of the radar; and<br />

5. Length and width of the SAR inagery.<br />

Ideally, the inagery will contain nany well-scattered control points<br />

on land whim can t:e located on an NI'S reference map. Fran the control nap, the<br />

latitude and longtitude of a control point are calculated and converted to real<br />

reference distances in the along-track and across-track directions. Similarly,<br />

measurements are taken fran the inagery in the along-track and across-track<br />

directions and converted to real inage distances.<br />

Using all the control points, a least-squares method is used to<br />

OOIlP1te unique scales and offsets in both the along-track and across-track<br />

directions. '!he result is two linear equations for the inagery, with reference<br />

values dependent upon the measured image values. In this manner, the location<br />

of any unknown point can t:e determined, provided one of the follOWing parameters<br />

is known:<br />

1. Target's location on an NTS nap;<br />

2. Target's latitude and longitude;<br />

3. Target's along-track and across-track real distances<br />

from the beginning of the flight; or<br />

4. Target's along-track and across-track inage distances.<br />

4. ICE MOl'Irn<br />

4.1 SLAR ANALYSIS<br />

Ice mtion during the first period (Decer.tJer 2nd to 6th) and the<br />

second period (December 14th to 16th) was determined by measuring the change in<br />

position of identifiable ice features. A set of approxinately 60 ice features<br />

was located on both pairs of inagery. Latitudes and longitudes of eam feature<br />

were calculated by neasuring the distance to the biO adjacent time r.arks, and<br />

interpolating exact positions using the INS data. '!his process assumed constant<br />

along-track and across-track scale and no INS drift.<br />

594


LARGE WINTER ICE MOVEMENTS<br />

IN THE NEARSHORE ALASKAN BEAUFORT SEA<br />

DAVID J. AGERTON SHELL OIL COMPANY HOUSTON, TEXAS, USA<br />

ABSTRACT<br />

Observations of a mid-winter storm in the Alaska Beaufort Sea and associated<br />

ice movements and deformations are discussed and <strong>com</strong>pared with simple mathematical<br />

force models. Landsat imagery and weather records from previous years are reviewed<br />

for indications of similar large ice movements.<br />

INTRODUCTION<br />

Ice movements are of interest because their rate, in part, determines<br />

structure loading. Also, movements create ice ridges, pile-ups, and leads which can<br />

load structures and disrupt over-the-ice logistics operations. The magnitude of ice<br />

movements determines, in part, the probability of ice features, such as multiyear<br />

floes, colliding with offshore structures in winter.<br />

In this paper, "large" movements are those exceeding 50 meters (m) per day.<br />

Some exceed 300 m per day. "Winter" and "nearshore" refer to movements which occur<br />

from December through May in waters up to 20 m deep.<br />

Past Investigations of Large Winter Ice Movements<br />

In protected, nearshore areas of the Beaufort Sea offshore of the North<br />

Slope of Alaska, winter ice movements are generally small and slow. The largest 30day<br />

excursion measured shoreward of the coastal barrier islands by industry in a<br />

three-year winter program was less than 4 m, and the fastest rate was about 1.5 m/hr<br />

(.001 ft/sec) [1]. In exposed areas, generally outside the barrier islands in water<br />

depths between 6 and 20 m, winter ice movements are larger, but are still constrained.<br />

In 20 m water depths, the largest measured excursion was about 70 m and the maximum<br />

rate was about 50 m/hr (0.05 ft/sec) [2]. These maximums are two to three orders of<br />

magnitude greater than their respective medians. In shallower water depths, the<br />

599


magnitudes and rates of ice movement decrease. In deeper water depths, they increase.<br />

At a location 24 km offshore, CRREL investigators measured about 1 km displacement<br />

during a moderate storm [3]. A correlation between large ice movements and the<br />

direction and sequence of storms in the nearshore Beaufort Sea has been observed [2].<br />

Although threshold windspeeds appeared to be required for large ice movements to<br />

occur, above windspeeds of 13 m/sec (20 knots), ice displacements do not correlate<br />

with local windspeeds.<br />

Shapiro [4] measured very rapid ice movements in the Chukchi Sea off<br />

Barrow in waters less than 20 m deep during an intense storm in late December 1973.<br />

The ice was 0.6 m thick. Onshore radar recorded ice feature position. Calculated<br />

ice movement rates were as high as 2.3 m/sec (7.6 ft/sec) in winds 26 m/sec (50<br />

knots), reportedly gusting to 52 m/sec. 2.3 m/sec is a much higher rate than observed<br />

in the Beaufort Sea. And, it is higher than might be expected. The arctic oceanographer's<br />

rule-of-thumb is that free-floating ice drifts at three percent of the wind<br />

speed. By this rule, the ice would have moved at 0.76 m/sec. The high-speed wind<br />

gusts and local ocean currents might have increased floe speed beyond that expected<br />

for the free-floating case. Considering it was December and the ice was probably<br />

<strong>com</strong>pact instead of free-floating, the high floe speeds are even more surprising.<br />

March 1979 Field Observations<br />

On March 17, 1979, a large extratropical cyclone developed in the Arctic<br />

Ocean northeast of the Beaufort Sea. The next day, it drifted southwestward and<br />

intensified. At Deadhorse airport, about 16 km inland, winds were 17 m/sec (33 knots)<br />

from 240 to 260°. The coastline is oriented at about 300°, so that a <strong>com</strong>ponent of<br />

the wind blew offshore. It is <strong>com</strong>mon knowledge that Eskimos stay off the North Slope<br />

ice during strong southwesterly winds because leads form and large ice movements<br />

typically occur.<br />

The storm coincided with several oil industry and government projects. So,<br />

an unusual amount of data was available to describe the storm and its effects. In<br />

addition to onshore coastal meteorological stations, an array of meteorological buoys<br />

was deployed in the Beaufort Sea and Arctic Ocean. The day preceding the storm, and<br />

again several days after the storm, NASA Lewis Research Center flew aerial photo and<br />

side-looking airborne radar (SLAR) missions in the area. Two days after the storm, a<br />

field team arrived to survey first-year ice features [5]. Refrozen leads, offset<br />

seismic roads, pressure and shear ridges, and a large ice pile-up appeared newly<br />

formed. The Landsat satellite passed over the area on March 12 and again on March<br />

20. Fortunately, both days were cloud free. In this same area, on March 29, Gulf<br />

R&D Co. had aerial photos made covering 500 km of flightline.<br />

600


shear ridge surveyed in March and seen on Landsat. It might have formed on November<br />

26, when a moderately intense easterly storm occurred with winds from gO-100°.<br />

The shear ridge struck an angle of 300°, so a <strong>com</strong>ponent of the wind was normal to it,<br />

but the major <strong>com</strong>ponent of the wind was parallel to it--just what would be expected<br />

for creation of a shear ridge.<br />

Unfortunately, where the ice was rough, the SLAR image return saturated the<br />

photographic paper and detail was obliterated. It was impossible to differentiate<br />

between areas of low, rough rubble and areas of high ridging. Ridges and narrow<br />

leads could also give indistinguishable returns. Consequently, although the SLAR<br />

imagery fixed the date of initial formation of a major offshore shear ridge, it was<br />

not helpful in determining the magnitude and direction of subsequent displacements,<br />

specifically those which occurred in March.<br />

Photography<br />

Aerial photos showed the magnitude, direction, and general location of<br />

nearshore ice movements. They showed patterns of ice failure. And, they established<br />

the location of ridges for a summer ice gouge survey by the USGS.<br />

principal flight line locations.<br />

Reference 11 shows<br />

Ice movements near Cross Island. NASA photos from March 16 showed straiqht<br />

seismic roads east of the island. A Gulf R&D Co. photo from March 29, showed these<br />

roads offset repeatedly by ice movements and numerous leads, changes attributed to<br />

the March 17 storm. Several large shear displacements were evident: 250 m at 130°<br />

and 110 m at 150°. The displacements occurred 2 to 3 km offshore in a sector between<br />

NNE and E of the island. Water depth here is roughly 14 m. Figure 3 is a line<br />

FIGURE 3. DRAWING OF SEISMIC ROAD DISPLACE­<br />

MENTS, LEADS, AND FAILURE PATTERN<br />

IN ICE EAST OF CROSS ISLAND,<br />

MARCH 1979.<br />

drawing tracing the offset seismic<br />

roads, shear failure lines, and<br />

ridged areas. It shows the wedgelike<br />

local ice failure pattern.<br />

Ice movements north<br />

of Narwhal Island. Figure 4 is a<br />

line drawing from Gulf photos<br />

offshore Narwhal Island [5] showing<br />

larger movements. Based on seismic<br />

road dispacement, ice moved eastward<br />

1200 m along a line oriented at<br />

300°. In the process, a 21 m high<br />

grounded ice pile-up formed, which<br />

the field team dubbed "Ice Mountain". Water depth here was 18 m. Further offshore,<br />

we know from Landsat that the ice moved 2500 m eastward. Closer to land, the ice was<br />

604


DISCUSSION OF MATHEMATICAL MODELS<br />

Grounded Ice Pile-ups<br />

The forces acting to create grounded ridges and pile-ups are generally<br />

thought to be low [12]. Most equations describing forces during ridging and pile-up<br />

result in predictions of forces less than 0.3 mega newtons/m (MN/m) (20 kips per<br />

linear foot). Vaudrey [5] estimated average forces during pile-up of Ice Mountain as<br />

being about 0.09 MN/m (6 kips/ft or 8 psi). He assumed that most of the work done<br />

during ridging was stored as gravitational potential energy, and the energy expended<br />

in friction and in creating new surfaces by fracture was small. Fracture energy loss<br />

in ridging appears negligable. But, friction forces may be of the same order of<br />

magnitude as gravity forces [12]. Kovacs and Sodhi [13] estimated the frictional<br />

force for an ice sheet being pushed up an inclined surface as a function of friction<br />

coefficient and geometry. Applied to Ice Mountain, their equation results in a<br />

calculated frictional force of 0.03 to 0.07 MN/m (2-5 kips/ft), depending on parameter<br />

assumptions. However, the frictional force along the shear boundary surface of<br />

the moving ice sheet was not estimated because the normal forces along the two<br />

surfaces shearing past one another were not known.<br />

Ice Dynamics<br />

Generally, the principal force moving ice is wind stress. The floe which<br />

created Ice Mountain appeared on Landsat to be about 1.6 km wide by 24 km long,<br />

assuming the western-most lead was a boundary. The forces which might have acted on<br />

this floe at a moment in time in the March storm were considered by estimating the<br />

inertial, drag, Coriolis, and deformation forces. Driving forces transmitted from<br />

adjacent floes were not considered.<br />

It was assumed that ice movement occurred during the 8-hour storm peak when<br />

winds exceeded 15 m/sec, that acceleration was sinusoidal, and that velocity peaked<br />

when half the 1200 m movement had occurred. This resulted in a total movement time<br />

of 4.4 hours and a peak velocity of 0.15 m/sec (0.5 ft/sec). This velocity was an<br />

order of magnitude greater than values measured in this area.<br />

Estimated coriolis, inertia, and water drag were 3 to 5 percent of wind<br />

drag on the ice. Thus, wind drag, deformation forces and driviriq forces from<br />

adjacent floes should about balance. Estimated wind stress on the floe was 37MN<br />

(8300 kips). The estimated pile-up and friction force for Ice Mountain was 13 MN<br />

(3000 kips) or about 35% of the wind drag. Photos showed many other areas of ice<br />

deformation,along the floe's southern boundary so that 13 MN should probably represent<br />

much less than 35% of the total driving forces. The floe was probably driven by<br />

the ice pack northwest of it as well as by local wind stress.<br />

606


Landsat Imagery 1973-1978:<br />

OTHER LARGE ICE MOVEMENTS<br />

Landsat imagery from 6 winters was examined for nearshore<br />

leads to see how often they occurred in the lease sale area. Landsat could not<br />

provide a <strong>com</strong>plete data sample because of low pass frequency and inability to image<br />

ice through cloud cover. The inferred ice movement and storm which appears most<br />

likely to have caused the leads by virtue of its date and wind direction are<br />

summarized for three examples below. A fourth example was discussed in reference 2.<br />

Identification of the associated storm may have been limited by poor wind data.<br />

March 14, 1973. The image is depicted by a line drawing in Figure 5. A<br />

refrozen lead 1 km wide occurred in 20 m of water 6 km N.E. of Cross Island. N.E. of<br />

Pole Island 7km, a 2 km wide lead extended into 17 m of water. A large area of pack<br />

ice had been displaced eastwards.<br />

and February 2.<br />

Moderate westerly storms occurred on January 29<br />

- .......... _--... _ 2O-fotIETERS _ ... ---:<br />

MARCH 14, 1973<br />

FIGURE 5. LINE DRAWING OF 3/14/73 LANDSAT IMAGE<br />

SHOWING PATTERN OF LEADS.<br />

KM<br />

March 10, 1974. An<br />

older, refrozen lead system<br />

was visible bordering on the<br />

20 m water depth contour.<br />

Estimated eastward displacement<br />

is 5 km at locations 8 km<br />

North of Pole Island and 25 km<br />

North of Pingok Island.<br />

Harrison Bay was free of leads<br />

in this and in other images.<br />

The winter of 1973-74 had the<br />

most intense storms of those<br />

examined: five westerly<br />

storms with winds exceeding 50<br />

knots. Storms preceeding<br />

March 10 occurred on January<br />

13 and January 27.<br />

April 19, 1975. A lead 1 km wide at the 20 m depth contour offshore<br />

Alaska Island was visible. Again, offshore areas west of Cross Island are lead free.<br />

A weak westerly storm occurred on April 14. It was preceeded by an intense easterly<br />

storm on March 28.<br />

SUMMARY<br />

Despite limitations in frequency and resolution, Landsat imagery has<br />

provided data on past ice movements in the nearshore Alaskan Beaufort Sea. Ice<br />

displacements of 1 to 5 km have been observed on Landsat images and aerial photos in<br />

607


R. S. Pritchard, Sr. Research Scientist<br />

M. D. Coon, Sr. Research Scientist<br />

CANADIAN BEAUFORT SEA ICE CHARACTERIZATION<br />

Flow Research Company<br />

Kent, Washington 98031 U.S.A.<br />

Abstract<br />

A model characterizing the Canadian Beaufort Sea ice cover in terms of fractions of<br />

coverage of open water (including new ice) and thin, flat and rubbled ice is presented.<br />

The minimum and maximum thicknesses that define each category vary in time<br />

to account for thermal growth, except for the fixed thickness separating open water<br />

(new ice) from thin ice. This characterization is intended to be part of a <strong>com</strong>plete<br />

ice dynamics model. Ice strength for a plasticity model is determined from the ice<br />

conditions. A specific function that redistributes only the thinnest ice available<br />

is introduced. This constitutive law can describe the essential physics of rafting<br />

and ridging processes and is so simple that it can be integrated analytically.<br />

Examples of model behavior with no deformation, isotropic opening, uniaxial closing<br />

and pure shearing allow the effects of several parameters on the response to be<br />

isolated. This result simplifies the determination of material constants when<br />

observed sea ice behavior is simulated. Future validation of model performance is<br />

suggested as actual data be<strong>com</strong>e available.<br />

Acknowledgement<br />

This work was supported by AES and DSS of the Canadian government and Dome Petroleum<br />

through a subcontract from Intera Environmental Consultants, Calgary, Alberta. We<br />

gratefully acknowledge the help of E. Leavitt, technical monitor, for a critical<br />

review of the work.<br />

609


<strong>com</strong>pression and pure shear. The response is acceptable for the material constants<br />

chosen. Ice strength depends strongly on the amounts of open water and thin ice<br />

present. As the fractions of open water and thin ice are depleted, the strength<br />

jumps. Values are 43 N/m for open water, 1.8 x 10 4 N/m for thin ice, 3.3 x 10 4 N/m<br />

for flat ice and infinity for rubbled ice. These values depend on other material<br />

parameters but are representative of values expected to allow accurate simulation of<br />

observed ice motions and deformations.<br />

While the behavior of this model is reasonable, no attempt has yet been made to<br />

<strong>com</strong>pare it with observed data. This critical step should be taken as soon as data<br />

are available. The model is rather simple to modify because the effects of various<br />

material constants can be isolated. This fact will allow constants to be tuned to<br />

match observations and the model's performance to be evaluated qualitatively.<br />

References<br />

1. Coon, M. 0., Maykut, G. A., Pritchard, R. 5., Rothrock, D. A., and Thorndike,<br />

A. S. (1974) "Modeling the Pack Ice as an Elastic-Plastic Material," in AIDJEX<br />

Bulletin 24, University of Washington, Seattle, pp. 1-105. ---<br />

2. Thorndike, A. 5., Rothrock, D. A., Maykut, G. A., and Colony, R. (1978) "The<br />

Thickness Distribution of Sea Ice," J. Geophysical Research, Vol. 80, No. 33,<br />

pp. 4101-4513.<br />

3. Rothrock, D. A. (1979) "Modeling Sea-Ice Features and Processes," l:.<br />

Glaciology, Vol. 24, No. 90, pp. 359-375.<br />

4. Coon, M. D. (1980) "A Review of AIDJEX Modeling," in Sea Ice Processes and<br />

Models, R. S. Pritchard (ed), University of Washington Press, Seattle, pp. 12-33.<br />

5. Pritchard, R. s. (1981) "Mechanical Behavior of Pack Ice," in Mechanical<br />

Behaviour of Structured Media, A.P.S. Selvadurai (ed), Elsevier, Amsterdam, to<br />

appear.<br />

6. Reimer, R., Pritchard, R., and Coon, M. (1980) "Consistent Reduction of Ice<br />

Thickness Distribution to a Few Categories," Flow Research Report No. 167, Flow<br />

Research Company, Kent, Washington.<br />

7. Nye, J. F. (1976) itA Coordinate System for Two-Dimensional Stress and<br />

Strain-Rate and its Application to the Deformation of Sea Ice," in AIDJEX<br />

Bulletin 33, University of Washington, Seattle, pp. 131-143. ----<br />

8. Rothrock, D. A. (1975) "The Energetics of the Plastic Deformation of Pack Ice<br />

by Ridging," J. Geophysical Research, Vol. 80, No. 33, pp. 4514-4519.<br />

9. Rothrock, D. A., and Hall, R. T. (1975) "Testing the Redistribution of Sea Ice<br />

Thickness from ERTS Photographs," AIDJEX Bulletin 29, University of Washington,<br />

Seattle, Washington, pp. 1-19.<br />

618


R. Colony<br />

A. S. Thorndike<br />

ABSTRACT<br />

SEA ICE STRAINS DURING 1979<br />

Polar Science Center<br />

University of Washington<br />

Seattle, Washington 98105<br />

A number of drifting data buoys were operational in the Arctic Basin from<br />

U.S.A.<br />

February through December 1979. Using the procedure of optimal interpolation, esti­<br />

mates of ice motion and gradients of ice motion were made at selected locations in<br />

the basin. Seasonal and regional patterns of strain are shown for both daily and<br />

longer periods of deformation. Strain and rotation statistics are <strong>com</strong>pared to<br />

measurements made during the Arctic Ice Dynamics Joint Experiment.<br />

INTRODUCTION<br />

Sea ice moves largely in response to stresses exerted by winds and ocean cur­<br />

rents. Even when these stresses vary smoothly in space the response of the ice pack<br />

can be uneven because of its granular character. The grains, single ice floes or<br />

groups of floes acting together and ranging in size from 10 to 10 5 meters, move<br />

rigidly with the only deformation occurring at the grain boundaries. A snapshot of<br />

the velocity field at any particular time would reveal a field which remained con­<br />

stant across a single grain (or varied linearly to account for rigid rotation of the<br />

grain) and which was discontinuous at the grain boundaries. At the grain boundaries<br />

energy is dissipated in building pressure ridges and in friction. When grains move<br />

apart new area of open water is exposed leading to vigorous exchange of heat between<br />

the ocean and atmosphere and rapid growth of new ice. Because of the range of grain<br />

sizes and the irregular geometry of the grain boundaries it is not feasible to<br />

resolve the deformation <strong>com</strong>pletely. Instead techniques are needed to characterize<br />

the deformational activity from limited measurements of ice motion. Toward this end<br />

we have taken recent measurements of ice motion which are widely separated in space<br />

and calculated a number of deformation indices from them.<br />

Evidence from earlier work suggests that such indices contain useful information<br />

about the behavior of the ice pack. In our analysis of the Arctic Ice Dynamics Joint<br />

619


DATA SET<br />

An array of automatic data buoys was established and maintained in the Arctic<br />

Basin as a part of the First GARP Global Experiment. The first buoys were deployed<br />

on 19 January 1979. The <strong>com</strong>plete array of about 15 buoys operated from the first of<br />

March 1979. Our analysis is restricted to the period 1 March - 31 December 1979.<br />

The measurement program is expected to continue for several more years. Objectives<br />

of the program are to provide measurements of surface atmospheric pressure and to<br />

define the large scale field of motion of sea ice. A data report [41, available from<br />

the authors, describes the measurement program, data processing procedure, and avail­<br />

able data sets.<br />

The deformation quantities are determined using techniques of optimal interpola­<br />

tion given in Gandin [51. The N measurements zi of displacement u i at points Xi with<br />

measurement errors E.<br />

1-<br />

The estimator is<br />

where the C1. i satisfy<br />

Here<br />

Zi - u i are used to estimate the displacement U at the point x.<br />

N<br />

E<br />

i=l<br />

1'. • U.U.<br />

1-J 1- J<br />

£.u.<br />

1- J<br />

N<br />

U = E C1. Z<br />

i i<br />

i=l<br />

o for i # j, and<br />

The correlation 1'ij between displacements at xi and Xj is taken to be a function of<br />

the separation 1'ij = R(lxi-Xjl). Similarly Sj is the correlation between displacement<br />

at x. and at X; B. = R(lx-x./). An early sketch of the correlation function R<br />

J J J<br />

was given in [11. For these calculations R(a) = q2 exp [-(a/Z)21 was used with q = 5 km<br />

and Z = 600 km. The above description for scalar quantities is readily extended to<br />

vectors, producing in the end estimates for the displacement <strong>com</strong>ponents U and V. The<br />

measured displacements zi were taken over one day intervals; zi = Xi(t + 1/2) -<br />

X.(t - 1/2) and x. = X.(t) where X. was the trajectory of the i th buoy.<br />

1- 1- 1- 1-<br />

(1)<br />

(2)<br />

621


Note that the matrix M = rr .. + a 2 o. ]-1 depends on the measurement positions<br />

L'LJ 'L.tJ<br />

xi but not on the position x where the interpolated value is desired. Thus the gra-<br />

dients of the estimated displacement can be found by <strong>com</strong>bining equations (1) and (2)<br />

and differentiating,<br />

N N ds.<br />

:E :E Mij zi a:i=l<br />

j=l<br />

We use the symbol U x for this quantity and u y ' V x ' Vy for the other derivatives of<br />

the estimated displacement field.<br />

The position measurements, obtained by satellite Doppler positioning techniques<br />

had essentially zero mean random errors with a standard deviation of about 300 meters ..<br />

This contributes to an interpolation error for the daily displacement which typically<br />

had standard deviation less than 500 meters at points within the region defined by<br />

the outer buoys. Extrapolations to regions beyond the outer buoys had much larger<br />

estimation errors. The nine grid points used in this study, Figure 2, were within<br />

the buoy array for the ten month period.<br />

STATISTICS OF DAILY STRAIN<br />

Figure 2. The grid points used<br />

in the strain study. This set of<br />

grid points was always inside the<br />

outer perimeter of FGGE buoys.<br />

The drift of the AIDJEX main camp<br />

is shown for May 1975 through<br />

April 1976.<br />

Spectral analysis of the velocity and velocity gradient time series shows little<br />

energy associated with periods of less than one day. This observation allows us to<br />

regard the daily displacement, u, and the instantaneous velocity, i, as nearly inter­<br />

changeable. Therefore in the description of strain which follows one can either think<br />

of the quantities, u x ' u y ' Vx and Vy as representing gradients of the estimated daily<br />

displacement (strain) which are dimensionless, or as gradients of the estimated<br />

622


- ':::>' -<br />

Figure 4. The deformation<br />

ellipses and total rotation angle<br />

for the selected grid points.<br />

The angle is measured clockwise<br />

from lines parallel with the date<br />

line, 180·. March 1979 through<br />

December 1979.<br />

We can interpret the deformation as points initially defining a circle of unit area<br />

being rotated through an arc and then mapped onto an ellipse. The area of the<br />

resulting ellipse indicates net convergence or divergence; the eccentricity of the<br />

ellipse is a measure of net shearing; and the orientation of the ellipse indicates<br />

the principal directions of the total deformation.<br />

The magnitude of the major and minor axes of the ellipses of Figure 4 can be <strong>com</strong>­<br />

pared to Figure 1, the AIDJEX array which accumulated strain for 360 days. Again we<br />

see similarities in the amount of total shear. A pattern of orientation of the ellip­<br />

ses is not clearly discernible. Perhaps there is a suggestion the major axis of the<br />

ellipse tends to be alligned with the coast. This may be due to either stress propa­<br />

gating out from the coast or the dynamic topography of the underlying ocean.<br />

The time evolution of the cumulative strain is seen in the time series of the<br />

major axis, minor axis, and the orientation of the major axis. The time series of the<br />

major and minor axes for grid point G, Figure 5, are qualitatively different from<br />

AIDJEX and some of the other grid points, in that the curves are not monotonic.<br />

Figure 5a clearly shows a prolonged extension (May through mid August) followed by a<br />

period of contraction (mid August through mid October) in that same direction, Figure<br />

5b.<br />

Figure 4 shows the ellipses G, H, and I to have similar configurations. However<br />

Figures 6a,b and 7a,b reveal marked differences in how they evolved to that final con­<br />

figuration. On time scales of tens of days the deformations at points roughly 500 km<br />

apart appear to be poorly correlated.<br />

As seen in the mean daily rotations, the region shows a net clockwise rotation.<br />

Time series of rotation show in general the steady evolution of the total rotation.<br />

The net divergence is small in most cases, and no pattern is suggested. In fact it<br />

appears that one cannot even determine the sign of the divergence for the entire region.<br />

625


DISCUSSION<br />

Our conclusion for the AIDJEX study that long term deformations were strongly<br />

organized over scales of 100-500 kilometers cannot, on the basis of these observations,<br />

be extended to much larger scales. Perhaps observations from subsequent years will<br />

confirm the pattern noted for the principal directions of the long term deformation.<br />

The results to date suggest the following tentative statements about the annual<br />

deformation:<br />

i) this part of the ice pack has a net clockwise rotation of about 40°-60° with<br />

the larger rotations appearing at the lower latitudes,<br />

ii) the long term deformation in this region is approximately non-divergent,<br />

iii) there is a net stretching of order 80% in one principal direction and con­<br />

traction of order 40% in the other.<br />

The general impression is that the deformations of regions 500 kilometers or more<br />

apart evolve more or less independently of each other, organized only weakly by some<br />

very large scale influence (such as the long term circulation of the atmosphere or<br />

ocean or the dynamic constraints imposed by the boundaries). This lack of organiza­<br />

tion is puzzling at first considering the well documented large scale organization<br />

of the velocity field itself consisting of a general clockwise circulation in this<br />

part of the basin. Daily velocities at different points in fact have strong positive<br />

correlations out to large distances (e.g., R(500 km) z 0.5). The paradox is explained<br />

by noting that the process of differentiation in going from displacement to deforma­<br />

tion inevitably accentuates the smaller length scale variations in the field of<br />

motion. This reduces the distance over which correlation is felt.<br />

Acknowledgment This work was supported by the National Oceanic and Atmospheric<br />

Administration Grant NA80-AA-D-00015, which was funded in part by the Global Atmospheric<br />

Research Program and the Office of Climate Dynamics, Division of Atmospheric<br />

Sciences and the Meteorology Program, Division of Polar Programs, of the National<br />

Science Foundation, and the Office of Naval Research, Arctic Programs.<br />

References<br />

[1] Thorndike, A. S. and R. Colony, 1980. Large-scale ice motion in the Beaufort<br />

Sea during AIDJEX, April 1975-April 1976, in Sea Ice Processes and ModeLs,<br />

R. S. Pritchard, Ed. University of Washington Press, Seattle, Washington (249-260).<br />

[2] Rothrock, D. A. and R. T. Hall, 1975. Testing the redistribution of sea ice<br />

thickness from ERTS photographs, AIDJEX BuLLetin. 29, July (1-19).<br />

[3] Rothrock, D. A., R. Colony and A. S. Thorndike, 1980. Testing pack ice constitutive<br />

laws with stress divergence measurements, in Sea Ice Processes and ModeLs,<br />

R. S. Pritchard, Ed. University of Washington Press, Seattle, Washington (102-112).<br />

[4] Thorndike, A. S. and R. Colony, 1980. Arctic Ocean Buoy Program, Data Report,<br />

19 January 1979 - 31 December 1979. Polar Science Center, University of Washington<br />

(131 pages).<br />

627


References, continued<br />

(5) Gandin, L. S., 1965. The Ob.iective Analysis of Meteorological FieUs.<br />

Leningrad, 1963, English translation, Israel Program for Scientific Translations.<br />

Jerusalem (242 pages).<br />

(6) Colony, R., 1978. Daily rate of strain of the AIDJEX manned triangle, AIDJEX<br />

Bulletin, 39, May (85-110).<br />

628


is equal to the product of the effective failure pressure of the ice by the diameter<br />

of the platform and by the thickness of the ice. The design of the structure is<br />

therefore governed by the value of ice failure pressure averaged over the full contact<br />

area. Our knowledge of this effective ice failure pressure over large areas will<br />

significantly affect the safety and cost of future production platforms.<br />

SCALE EFFECT ON ICE STRENGTH<br />

It has been known for a long time that the measured strength of ice is some function<br />

of the size of the sample. (Weeks & Assur 1969, Metge 1977, Gold 1978). In fact,<br />

Weeks and Assur (1969) state that, "an understanding of the scale effect in ice testing<br />

is essential before a thorough scientific basis can be developed for the utiliza­<br />

tion of small scale testing in engineering design problems". Their statement applies<br />

just as well today as it did 12 years ago.<br />

In the case of production platforms, the ice failure area is typically 3000 m 2 and<br />

the failure pressure must be deduced from indentation tests involving a few square<br />

2<br />

centimetres or at the very most 3 m • Extrapolating from 3 to 3000 m2 is dubious<br />

indeed, especially since the scale effect phenomenon is not yet well understood.<br />

As an example of how widely opinions may vary regarding scale effect, consider the<br />

following relationships between the failure pressure P, the ice thickness h and the<br />

width of the failure area D: Weeks and Assur (1969) based on data by Butiagin<br />

thought that Po(D-O. S h-O. S • Hirayama et al (1974) show that PeCD-O. S h+O· 1 • More<br />

extensive tests by Saeki et al (1977) indicate hCD-o· S hO.O. While Gold (1978)<br />

found that PI(D-0. 4 hO.O seems to fit data from a variety of sources. More recently,<br />

Kry (1979) has shown that the effect of D (in terms of number of zones). on P decreased<br />

when D be<strong>com</strong>es large and eventually that PO( DO. 0.<br />

Knowing how important P is to the design of production platforms, and knowing the<br />

amount of uncertainty in methods of extrapolating small scale tests to large scale<br />

ice failure, it be<strong>com</strong>es obviously very attractive to try and measure large scale<br />

failure pressures directly in the field.<br />

STATE OF THE ART IN ICE FORCE MEASUREMENTS<br />

Many measurements of ice forces on narrow structures such as bridge piers (Neill<br />

1976), lighthouses and drilling platform legs (Blenkarn 1970) have been made. The<br />

measured ice failure pressures for narrow structures range from 1 to 3 MPa.<br />

To date, however there has been no direct measurement of ice forces on wide struc­<br />

tures. In order to evaluate ice forces on their artificial islands, oil <strong>com</strong>panies in<br />

Canada and Alaska have used ice stress sensors embedded in the ice sheet surrounding<br />

630


the artificial island, (Metge et al 1975). Using the ice stress at a few points and<br />

an elastic model of stresses in a plate around a fixed object, one can derive an<br />

estimate of the total ice forces. Very few recorded stress events (if any) actually<br />

involved ice failure because most island locations have been in landfast ice. A<br />

measurement of ice stress during ice failure is reported by Sackinger 1979. The<br />

disadvantages of such measurement methods are:<br />

-the questionable accuracy of the stress sensors<br />

-the possible additional error introduced in deriving total ice forces from<br />

local ice pressures.<br />

THE HANS ISLAND PROJECT<br />

During 1979, Dome Petroleum initiated a project with the aim of measuring full scale<br />

multi-year ice forces on stationary obstructions (Metge 1979). Several different<br />

possibilities for such measurements were reviewed, such as:<br />

-instrumented test structure<br />

-giant "nutcracker" type of device using hydraulic jacks to fail the ice<br />

over large areas<br />

-instrumentation of natural rocks or islands<br />

-instrumentation of the ice failing against natural rocks or islands<br />

As a result of this study, it was found that it might be possible to measure large<br />

scale ice forces by measuring the deceleration of large fast moving floes as they<br />

impact a natural obstruction. At the same time, a survey was made of all the rocks<br />

and small islands in the Canadian Arctic which might be used to measure full scale<br />

multi-year ice forces (Roche 1979).<br />

After air photo surveys of the most likely locations, it was realized that Hans<br />

Island, located in the middle of Kennedy Channel between Greenland and Ellesmere<br />

Island at 81° latitude North, provided an ideal platform from which to make measurements<br />

of ice floe decelerations during impact and therefore determine large scale ice<br />

forces. During August 1980, a field party of five stayed on Hans Island for three<br />

weeks and measured the decelerations of floes up to 6 km in diameter as they impacted<br />

the island at speeds up to 0.6 m/s.<br />

2. HANS ISLAND FOR ICE RESEARCH<br />

Hans Island (Figure 1) is a location unique in the world and extremely well suited to<br />

determining ice design criteria for Beaufort Sea platforms. It is a steep island, 1<br />

km in diameter and 150 m high which stands in the middle of Kennedy Channel. The ice<br />

conditions in Kennedy Channel in July and August are similar to the worst ice condi-<br />

631


tions expected in the Beaufort Sea during a summer polar pack invasion, i.e., large<br />

multi-year floes up to 20 km in diameter, with average ice thicknesses up to 7 m<br />

moving down the channel and impacting Hans Island at speeds over 1 mls (due to high<br />

winds and currents in the channel).<br />

Even ice islands have been known to impact Hans Island as when ice island WH5, about<br />

8 x 20 km, became stuck between Hans Island and Ellesmere Island for several months<br />

(Nutt 1966).<br />

The highest ever recorded ice pile-up (25 m) was recorded at Hans Island in 1974 by<br />

M. Dunbar, when a 7 m thick 16 km by 10 km multi-year floe, again became stuck<br />

against Hans Island. The floe was held up by the island and kept moving back and<br />

forth with the tide, grinding against the island for over a month (Kovacs 1979).<br />

The disadvantages of Hans Island are: the difficult logistics, the very short season<br />

(the ice breaks-up about July 27 on the average and the weather closes in a month<br />

later), and the high frequency of fog.<br />

3. SOME QUALITATIVE OBSERVATIONS DURING ICE IMPACT<br />

During the summer of 1980, eight impacts of multi-year floes larger than 10 km 2 on<br />

Hans Island were recorded over the 21 day field work.<br />

In most cases the mode of ice failure at the interaction zone could be called a<br />

"crushing" mode, with very little evidence of any flexural failures. The failed ice<br />

showed evidence of a <strong>com</strong>bination of "flaking" and "crushing".<br />

In some cases the large floe was preceeded by a "cushion" of smaller floes which were<br />

squeezed against the island and absorbed the impact by ridging and rafting.<br />

In other cases the large floes were split by the island upon impact. This occurred<br />

in particular when the floe was made-up of several smaller multi-year pieces in a<br />

matrix of first year ice, a not un<strong>com</strong>mon occurrence.<br />

4. IMPACT FORCE MEASUREMENTS: METHODOLOGY<br />

GENERAL PRINCIPLES:<br />

When a large independent ice floe collides with a rigid obstruction, it decelerates<br />

rapidly due to the force required for ice failure. The interaction dissipates the<br />

kinetic energy of the floe through work done to fail the ice. Wind drag, current<br />

drag and Coriolis force may also contribute to the interaction. It is usually assumed<br />

that the energy of deformation inside the ice floe is negligible <strong>com</strong>pared to<br />

632


points, A and B. After the fact, the location of G can be determined from the shape<br />

of the floe and its thickness distribution and the distances AG and BG can be meas­<br />

ured. The acceleration of G can then be calculated from the two equations:<br />

... ... -t ...<br />

a G = a A + w AG<br />

... ... -t ...<br />

a G - a G + w • BG<br />

Note here that while the two accelerations a A and a B are required to calculate a G ,<br />

they, as an added benefit, give the value of w.<br />

In summary, the measurements required to apply equation (l) are: area of floe,<br />

average ice thickness, representative ice bottom roughness, shape and thickness<br />

distribution of the floe to determine tbe location of G, and either acceleration of<br />

center of mass G, or acceleration of A (magnitude and direction), acceleration of B<br />

(magnitude and direction), direction of AB.<br />

It is clear that there would be a great advantage in determining the location of G<br />

before the impact and measuring the acceleration of G only.<br />

In the special case where w stays constant throughout the impact, the acceleration is<br />

the same for any point over the floe. This happens in particular when a floe with no<br />

initial angular velocity hits the island "head on", and <strong>com</strong>es to a <strong>com</strong>plete stop. In<br />

this case, one acceleration measurement is sufficient to calculate the force.<br />

b) The application of equation (2) requires knowledge of: the magnitude of the<br />

velocity of the centre of mass of the floe at two given instants (it's direction is<br />

not necessary), the angular velocity of the floe at those two instants, the mass and<br />

moment of inertia of the floe, the displacement of the center of mass between the two<br />

instants.<br />

Note that this method only gives the average magnitude of the <strong>com</strong>ponent of the resul­<br />

tant force which is parallel to the displacement ds, it does not give either the<br />

magnitude of the force or its direction.<br />

Equation (2) is therefore generally of limited use, however, it can be very useful,<br />

in the special case where a floe with no initial angular velocity <strong>com</strong>es to a <strong>com</strong>plete<br />

stop against the island. In this case, equation (2) reduces to I: F • d; a 1/2 MV 2 •<br />

... ... 0<br />

Furthermore, in general in this case, F is parallel to ds, and, if the total displacement<br />

of G between the time of impact and the time of stopping is S , the total<br />

penetration, equation 2 then simplifies to:<br />

634<br />

F • S a 1/2 MV 2<br />

o<br />

(4)


sumably is obtained by multiplying the "ice failure pressure" corresponding to the<br />

worst possible "failure mode" and the worst probable ice, by an adequate safety factor,<br />

or "load factor".<br />

Almost always, the definition of "ice pressure" has implicitly corresponded to the<br />

case of a head-on collision without rotation; i.e. to the case where the force, F,<br />

the velocity of the centre of mass V G , the velocity of ice at the point of contact<br />

VA' and the normal to the contact area n, are all on the same line. The 'effective ice<br />

pressure is then simply defined as the ice force divided by the width of the contact<br />

area and by the ice thickness. However, in a more general case, illustrated in<br />

Figure 2, the directions of V G , VA' F and n are all different, and they all change<br />

during the impact. Defining the ice pressure in this case is more difficult: we<br />

know the force F and the ice thickness, but what width of contact area W should we<br />

use? There are four conceivable definitions of the effective ice pressure.<br />

1. Using the width in the direction of the ice force, F, PI - _----!F'--_<br />

h W cos a<br />

2. Using the width in the direction of motion of the ice which is failing, VA<br />

P =<br />

2<br />

3. Using the maximum width P 3<br />

F cos (0 - a)<br />

h W cos 0<br />

po cos a<br />

hW<br />

(wi th a friction factor f = tan S)<br />

As an example, using co = 15°, a - 30°, 0 = 60°, would give: PI<br />

1.73 F/wh, P 3 = 0.87 F/wh. Which one is correct?<br />

1.15 F/Wh, P 2<br />

The word "pressure", i.e. "the force per unit area exerted by a fluid", implies that<br />

the ice pressure should be the force per unit area in the direction normal to the<br />

contact surface.<br />

In a more practical way, to calculate the overall required lateral resistance of a<br />

structure, the engineer needs the design effective pressure P, normal to the face of<br />

the structure; he then calculates a maximum normal force Fn = P x Wand if nece­<br />

ssary the maximum shear force as F s = P x W x f where f is a sui table friction<br />

factor. Therefore definition 3 is most appropriate, i.e., if during the measurement<br />

of an impact, the ice force F and its angle a with the contact zone are known then<br />

the effective friction factor is f = tan a and the effective ice pressure is:<br />

P = F cos S!Wh<br />

Note that the effective friction factor during crushing may not be the ice/structure<br />

friction factor. If the sliding is occurring within the crushed ice, the friction<br />

factor f would be closer to the "internal friction" of crushed ice when it is consid-<br />

636


C. Theodolites<br />

If the impact lasts long enough, it is possible to use two theodolites mounted on top<br />

of the island. Two easily identifiable points on the approaching floe can be select­<br />

ed and the azimuth and declination of the points recorded as a function of time<br />

during the impact. If enough measurements are made, the motion of the centre of mass<br />

of the floe can be calculated as well as its deceleration.<br />

The deceleration obtained in this case is the average deceleration over the time<br />

required for three readings of azimuth and declination.<br />

D. Photogrammetry<br />

Photographs taken at regular intervals during the impact can also provide a measure<br />

of the deceleration. The photographs can be taken from the island itself or from a<br />

helicopter hovering over the contact zone. In this way the photographs provide a<br />

record not only of the translation and rotation of the floe but also of the ice<br />

failure mode and the width and shape of the failure zone.<br />

After the impact, the following measurements can be made in order to characterize the<br />

ice floe.<br />

a) Floe size - Using photographic or surveying techniques.<br />

b) Ice thickness - Three methods were used during the 1980 Hans Island project:<br />

direct measurements at augered holes, surveys of freeboard with rod and level, and<br />

using an impulse radar system mounted on the helicopter.<br />

c) Ice strength - The FENCO borehole jack system was used to provide an index of the<br />

ice strength, in order to be able to apply the data to other locations and different<br />

ice. This method provides a "confined crushing strength" and a modulus of elasticity.<br />

d) Crystallography, salinity, temperature, wind velocity and direction as well as<br />

current velocity and direction should also be recorded.<br />

6. EFFECTS OF WIDTH AND THICKNESS ON ICE PRESSURE<br />

The effect of size on the effective ice pressure was discussed briefly in the introduction.<br />

It was shown that the "size effect" is not yet well understood and that<br />

there are many varied opinions on the effects of the width D of the failure area and<br />

the ice thickness h.<br />

The following discussion is an attempt at better understanding this scale effect and<br />

638


REFERENCES:<br />

Blenkarn, K.A. (1970), "Measurements And Analysis Of Ice Forces On Cook Inlet Struc­<br />

tures", Offshore Technology Conference, paper 1261, Houston, U.S.A., Vol. II, pp 365-<br />

380.<br />

Gold, L. (1978), "Ice Pressures And Bearing Capacity", Geotechnical Engineering For<br />

Cold Regions, Edited by Andersland and Anderson, McGraw-Hill, ISBN 0-07-001615-1.<br />

Hirayama, et al (1974), "An Investigation Of Ice Forces On Vertical Structures",<br />

University Of Iowa, Institute Of Hydraulic Research, Report 158, Iowa City.<br />

Kovacs, A., and Sodhi, D.S. (1979), "Shore Ice Pile-Up And Ride-Up", Workshop On<br />

Problems Of The Seasonal Ice Zone. Naval postgraduate school, Monterey California,<br />

February 26-March I, 1979.<br />

Kry, R. (1977), "Implications Of Structure Width For Design Ice Forces", Int. Union<br />

Of Theoretical And Applied Mechanics, Symposium on the Physics of Ice Mechanics Of<br />

Ice, Copenhagen, August 6-10.<br />

Metge, M. et aI, (1975), "On Recording Stresses In Ice", <strong>Proceedings</strong> of 3rd Interna­<br />

tional Symposim of IAHR On Ice Problems, Hanover, U.S.A., pp 459-468.<br />

Metge, M. (1977), "Recent Field Testing Program", Workshop On The Mechanical Proper­<br />

ties Of Ice. NRC Technical Memorandum Number 121.<br />

Metge, M. (1979), "Full Scale Ice Force Measurements", Report For Dome Petroleum<br />

Ltd., August 12, 1979 (proprietary).<br />

Michel, B. (1978), "Ice Mechanics", Les Presses de I 'Universite Laval Quebec, ISBN<br />

0-7746-6896-8.<br />

Neill, C.R. (1976), "Dynamic Ice Forces On Piers And Piles". An assessment of design<br />

guidelines in the light of recent research. Canadian Journal Of Civil Engineering,<br />

Vol 3, pp 305-341.<br />

Nutt, D.C., (1966), "The Drift Of Ice Island WHS", ARCTIC 9 (3), pp 244-262.<br />

Roche, C. (1979), "A Selection Of Islands For Use As A Test Platform", Report by C­<br />

Core for Dome Petroleum Ltd.<br />

Sackinger, W., et al (1979), "Ice Stress Near Grounded Structures", <strong>Proceedings</strong> Of<br />

5th International Conference on POAC 79, Vol I, pp 57-73.<br />

Saeki, H., Hamanaka, K., Ozaki, A. (1977), "Experimental Study Of The Ice Forces On<br />

A Pile", proceedings of POAC Conference 1977, pp 695-706.<br />

Weeks, W. and Assur, A., (1969), "Fracture Of Lake And Sea Ice", CRREL Research<br />

Report 269.<br />

640


642<br />

a)<br />

NUMBER OF INDEPENDENT ZONES<br />

b)<br />

c)<br />

n • D/4h<br />

Pi • f(h,D), WITH D· 4h ... Pj. flh ONLY)<br />

FIGURE 3. THEORY OF SCALE EFFECT ON ICE PRESSURE ON WIDE<br />

STRUCTURES


ABSTRACT<br />

PROBABILITY DISTRIBUTIONS FOR STRUCTURE LOADING<br />

J. D. Wheeler<br />

Research Advisor<br />

BY MULTIYEAR ICE FLOES<br />

Exxon Production Research Company<br />

Houston, Texas<br />

Possible encounter with multiyear ice floes must be allowed for in the design of<br />

bottom-founded offshore structures in the arctic. Given a knowledge of sea-ice<br />

mechanics and of the failure mode imposed by structure geometry on a multiyear<br />

floe, there will be uncertainty as to the diameter, thickness and number of such<br />

floes that may impact a structure in different seasons of the year. This paper<br />

describes procedures for calculating probability distributions for multiyear ice<br />

loads on a structure during open-pack conditions (e.g., breakup, summer invasions,<br />

seasonal pack) and consolidated-pack conditions (winter). Information on floe<br />

diameters and thickness, areal ice coverage, ice movement and ice strength are<br />

<strong>com</strong>bined in a Monte Carlo calculation to develop the probability of exceedance<br />

(risk) versus loading in each season. Variation of ice thickness through the year<br />

and variation in point of contact between structure and floe are allowed for.<br />

Sensitivity of results to several types of environmental data is examined using<br />

reasonable but fictitious values for environmental parameters.<br />

Introduction<br />

In the design of offshore structures for arctic service, consideration must be<br />

given to possible encounter with multiyear ice floes. In this paper, Monte Carlo<br />

calculation procedures are outlined for estimation of load-risk curves for rapidice-movement<br />

conditions (summer) and for limited-ice-movement conditions (wintertime<br />

fast ice). The calculation procedures provide a framework for converting<br />

field observations on multiyear ice into probability distributions for ice loading<br />

and for evaluating the sensitivity of design-level loads to various input items.<br />

Distribution of Multiyear-Floe Forces in Summer<br />

"Summer" is defined here as the period from June 1 through November 1, which includes<br />

breakup, gross open-water season and freezeup. Summer ice coverage and multiyear<br />

fraction can be gotten, for example, from ice charts for a location of interest.<br />

Such coverage data is used in the calculation procedure to be described by assuming<br />

rough equality between the areal fraction of sea surface that is covered by multiyear<br />

ice and the fraction of total distance moved by a partial ice cover that brings<br />

multiyear ice against a fixed structure [6]. This assumption together with an<br />

estimate of ice-movement rate, permits translation of partial coverage by multiyear<br />

ice into an equivalent distance of multiyear ice that will impact a fixed structure.<br />

To implement the assumption, it is necessary to specify time series through the 643


MOVE FLOE<br />

PAST STRUC. 1----=:::::1:------1<br />

FIGURE 2. SUMMER CALCULATIONS<br />

DATA<br />

__ MYR. COVER<br />

DRIFT RATE<br />

-- FLOE THICKNESS<br />

IN WINTER PACK<br />

-- MAX. THICKNESS I<br />

-- FLOE DIAMETERS I<br />

__ MELT AND<br />

FREEZE RATES<br />

crushing strength against the structure. Following this period, the stress developed<br />

is the lesser of those from edge-failure or splitting [3]. Thickness,<br />

stress and structure diameter (D) give structure load. Another floe failure will<br />

not occur until the present floe is moved past the structure by ice motion. If<br />

this movement consumes the movement estimate for the current week, time is shifted<br />

forward one week; if not, another floe is sampled.<br />

The procedure of Figure 2 is continued until <strong>com</strong>pletion of a summer. The maximum<br />

force exerted on the structure during the entire summer is stored. The entire<br />

procedure is then repeated for 1000 summers. This population is histogrammed to<br />

determine the distribution of maximum multiyear loading for a randomly chosen<br />

summer. The development of a population of maximum forces for an entire summer<br />

amounts to an integration over the three distributed quantities (floe diameter,<br />

thickness and impact point on the structure), under the constraints that the sum of<br />

floe diameters in any week match the ice movement specified for the week and that<br />

floe thickness in any week be consistent with both the thickness distribution for<br />

the pack (Figure 3), a melting rate derived from air-temperature measurements and<br />

location water depth. It is these time-varying constraints that distinguish the<br />

procedure in Figure 2 from that used for multiyear ridges in [6]. In the ridgeloading<br />

work, no allowance was made for variation through the time period of<br />

interest (one summer) or statistics associated with individual ridge failures.<br />

645


Distribution of Multiyear-Floe Forces in Winter<br />

Winter is defined for calculations here as the period November through May. As<br />

with summer calculations, it is necessary to specify an areal coverage of multiyear<br />

ice and an ice-movement rate for each week. To make use of values input for multiyear<br />

coverage and movement rate, the assumption is again invoked that there will be<br />

rough equality between the areal fraction of multiyear ice and the fraction of<br />

multiyear ice along any linear path across the ice [6]. Intuitively, this assumption<br />

seems most likely to be satisfied when the multiyear ice is uniformly<br />

distributed over the area to which the specified fractional coverage is taken to<br />

apply. A detailed instance of uniform areal distribution of multiyear floes would<br />

be for each multiyear floe to have connected to it an area that is free of multiyear<br />

floes, in such a way that the areal percent of multiyear coverage is maintained<br />

in the vicinity of every floe. In summer, the floe-free area around each floe<br />

would contain water and vestigial annual ice. In winter, the floe-free area would<br />

consist of annual ice. The size of the floe-free area would increase with floe<br />

diameter. This picture does not preclude multiyear floes being in contact. There<br />

can be clusters of 2-4 multiyear floes and arbitrarily long lines of multiyear<br />

floes. The picture is somewhat like a thin section of cellular tissue, with<br />

multiyear floes being the cell nuclei and the floe-free area the cytoplasm, the<br />

whole cell being contained within a flexible membrane. In such a picture, the<br />

distance between floes can vary widely for any specified floe-size distribution,<br />

but the variation is constrained by the requirement that areal coverage be preserved<br />

"in the small"; i.e., in the vicinity of each floe.<br />

The above picture of detailed uniformity of multiyear-floe areal distribution<br />

brings consideration of floe spacing into the calculation procedure to be described<br />

below. Such consideration permits allowance for the fact that, in wintertime fast<br />

ice, the total movement and the multiyear coverage are not great enough for a<br />

structure to be contacted by a multiyear floe in every week, or even in every<br />

winter. In the summer calculation, it is tacitly assumed that ice movement in any<br />

week is always sufficient for multiyear floes to contact a structure, if such floes<br />

are present. The relatively limited movement in wintertime fast ice also invites<br />

consideration of percent-coverage values that are appropriate over areas that may<br />

contain several multiyear floes but are <strong>com</strong>parable in size to a total winter's<br />

movement. Historical information on multiyear ice coverage seems primarily to<br />

consist of visual estimates by trained ice observers from airplanes or ships. The<br />

area to which the visual estimates apply has dimensions on the order of miles, a<br />

region much larger than that swept out by a winter's movement in the fast ice. Any<br />

localized concentrations of multiyear ice are not recorded. Aerial photography of<br />

the fast ice indicates that it is not infrequent for multiyear ice to occur in<br />

patches, as indicated schematically in Figure 4. The floes in these patches could<br />

in some cases be fragments of larger floes that, weakened by summer melting, broke<br />

apart during the fall storms that moved them into the fast-ice zone. In any case,<br />

it is possible to define as a "multiyear area" the photographed area between the<br />

start and finish of such a multiyear patch. The multiyear coverage within such a<br />

multiyear area of flightline can be determined with reasonable accuracy to give a<br />

value of multiyear coverage that applies over a relatively small area in the waterdepth<br />

range spanned by the photography. Outside the multiyear areas, the ice is<br />

entirely first-year. A localised estimate of the chance that there will be a<br />

multiyear hazard in winter can be gotten from the ratio of total multiyear area to<br />

photographed area. Coverage values in multiyear areas of the kind defined in<br />

Figure 4 seem more appropriate for quantification of the mUltiyear-ice hazard in<br />

wintertime fast ice than visual averages over broad areas.<br />

The wintertime calculation flow for a specified location is given in Figure 5.<br />

Winter too is marched through, week by week. At the beginning of each week.a multiyear<br />

floe is sampled. Its thickness in the current week is gotten by sampllng<br />

647


a FLOGAP of ice past a structure to determine whether or not there is structurefloe<br />

contact in the current week. The average daily rate does not determine the<br />

rate of loading when structure-floe contact is made and so does not determine the<br />

crushing strength that the floe will exert against the structure. For a given<br />

hourly movement rate, ice strength is determined from laboratory and field data on<br />

strength variation with strain rate, as outlined in [4]. The strength determination<br />

is made for both annual (A) and multiyear (M) ice. The fraction of structure<br />

diameter that is loaded by each of these two ice types is determined by random<br />

sampling of the point along the structure diameter that is impacted by the center<br />

of the multiyear floe. For the force calculaton, strain rate and load are determined<br />

from the following two expressions.<br />

Strain rate = Ice velocity/(2Df D )<br />

Force = fcICx(DfD)t<br />

where D is structure diameter, fD is the factor by which grounded rubble around<br />

the island increases its effective diameter, fc is a contact factor [4], I is an<br />

indentation factor, C x is the unconfined <strong>com</strong>pressive strength of the ice and t is<br />

the ice thickness. Edge or splitting failure [3] is not considered in the winter<br />

calculations because ice floes are assumed to be confined in winter by the annual<br />

ice sheet. This should be conservative treatment of early winter, when the annual<br />

ice is thin. Following the force calculation, the ice is moved the remaining<br />

distance assigned to the current week, with time advance and/or selection of<br />

another FLOGAP as discussed above. As with the "summer" calculations, the maximum<br />

loading experienced by the specified structure in each winter season is stored and<br />

the calculation repeated for several thousand winters to obtain a population of<br />

maximum annual winter loads.<br />

Example Calculations<br />

Calculations indicating sensitivity or results to input quantities were run for<br />

schematic summer and winter cases. For summer, a structure in 60 feet of water was<br />

considered. The distance of multiyear ice moving past the structure in every year<br />

was gotten for each summer week using the ice-coverage and movement values from<br />

Figure 1 plus the assumption that 50% of the ice is multiyear during an invasion<br />

and 12.5% is multiyear otherwise. Floe diameters and thicknesses were sampled from<br />

Figure 3. Maximum ice thickness in the winter pack was specified to be 16 feet.<br />

The procedure of [3] was used to determine the mode of failure and stress imposed<br />

for each floe impacting the structure. The contact factor[f c in Eq. (2)] was set<br />

at 0.6 and indentation factor(I) at 3.0 [4]. Grounded rubble was assumed to be<br />

absent in the summer period. This summertime base case gave the solid line in<br />

Figure 7 for probability of exceedance versus normalized maximum summer load.<br />

Structure loads have been normalized by the product CxD from Eq.(2) to give units<br />

in Figure 7 of square inches per foot of structure diameter. Sensitivity to the<br />

product of ice coverage and ice movement was considered by arbitrarily dividing<br />

this product by 100 for each summer week. This gives the dashed curve(triangles)<br />

below the base-case curve in Figure 7. The reduction in ice movement makes little<br />

difference in the annual-risk range likely for design (0.1-0.01). Increasing the<br />

assumed maximum thickness of multiyear ice from 16 to 24 feet gives a corresponding<br />

50% increase in load for the 0.1-0.01 range of annual risk. This correspondence<br />

plus the steepness of the curves in Figure 7 indicate that the calculation procedure<br />

of Figure 2 gives loads associated with the maximum thickness of multiyear ice that<br />

can reach the structure. This maximum will be the lesser of water depth and the<br />

maximum multiyear thickness assumed to exist in the pack ice.<br />

650<br />

(1)<br />

(2)


Change in Calculation Input Change in Load at Specified Annual Risk<br />

Risk = 0.1 Risk = 0.01<br />

KIPL1000 % KIPL1000 %<br />

No change = Base Case 55 0 78 0<br />

Increase hourly rate x 10 107 95 151 95<br />

MYr cover = 0.375 66 20 94 20<br />

MYr cover = 0.125 in<br />

10% of years 51 - 7 73 - 6<br />

Median Floe diam. = 250 ft. 53 - 3 70 -10<br />

=1000 ft. 57 3 82 5<br />

MYr strength = 1.2 annual 57 3 84 8<br />

Increase max. myr thickness<br />

to 24 ft. 61 11 99 27<br />

The 95% increase in load with a ten-fold increase in ice velocity follows from<br />

Eqs.(l) and (3);10 raised to the 0.291 power is 1.95. The large increase indicates<br />

the need in these calculations for accurate determination of the relation between<br />

stress and strain rate and of ice movement rates in winter. The 20% load increase<br />

with a three-fold increase in multiyear ice coverage also stands out in the tabulation.<br />

However. available information indicates that 37.5% multiyear coverage in<br />

wintertime fast ice is rare. which reduces the significance of the load increase<br />

shown. The load-risk curve for 37.5% multiyear cover is plotted in Figure 8. The<br />

fourth table entry was obtained by assuming 1/8 multiyear coverage near the<br />

structure in only 10% of the winters. The coverage for other winters was uniformly<br />

distributed between zero and 1/8. The load-risk curve for this case is also<br />

plotted in Figure 8. The median floe diameter in Figure 3 is about 500 feet. Use<br />

of distributions parallel to the one in Figure 3 through median values of 250 and<br />

1000 feet changes loads in the 1-10% risk range by no more than 10%. Increasing<br />

multiyear strength by 9%. to 1.2 times the annual strength. gives essentially the<br />

same percentage increase in load at 0.01 annual risk but a lower increase in load<br />

at 0.1 annual risk. This is consistent with the low-risk events occurring when<br />

the structure diameter is almost <strong>com</strong>pletely spanned by a multiyear floe. Similarly<br />

a 50% increase in maximum allowable floe thickness to 24 feet has greater effect<br />

on the low-risk loading events. which are associated with a larger cross-section<br />

of multiyear ice impinging on the structure.<br />

REFERENCES<br />

1. Herbert. W .• Geographical J .• 136(4).511533. December. 1970<br />

2. Kniskern. F .. E .• Potcsky. G.J. 7'Frost Degree Day. Related Ice Thickness ..• ". Tech.<br />

Rpt. TR-60. Oceanogr. Prediction Div .• US Naval Oceanogr. Office. July. 1965<br />

3. Ralston. T.D .• "Plastic Limit Analysis of Ice-Splitting Failure".POAC 81 Preprints.<br />

4. Wang. V.S •• "Tech. Seminar on Alaskan Beaufort Sea Gravel Island Design".Exxon<br />

Company. U.S.A •• October 18. 1979<br />

5. Weeks. W.F .• et al."Characterization of Surface Roughness and Floe Geometry ••. ".<br />

AIDJEX Symp. on Sea Ice Processes .•.• Preprints. Vol.II.September 6-9. 1977<br />

6. Wheeler. J.D .• "Probabilistic Force Calculations for Structures in Ice-Covered<br />

Seas". <strong>Proceedings</strong>. POAC 79. p.1111-26. Vol.II. Trondheim. August. 1979<br />

652


A. B. CAmmaert,<br />

Manager, Engineering Studies<br />

G. P. Tsinker,<br />

Lead Engineer<br />

ABSTRACT<br />

IMPACT OF LARGE ICE FLOES AND<br />

ICEBERGS ON MARINE STRUCTURES<br />

Acres-Santa Fe Incorporated<br />

Calgary<br />

Acres Consulting Services Limited<br />

Niagara Falls<br />

In recent engineering studies it has been necessary to calculate the<br />

impact of large ice features on various types of marine structures.<br />

A simple approach was developed which relates kinetic energy dissi­<br />

pation to progressive crushing of the ice.<br />

When a large ice floe or iceberg collides with a massive structure<br />

the contact zone will fail by crushing, and will increase in size as<br />

the resisting force is steadily increased. The kinetic energy is<br />

then decreased until an equilibrium point is reached.<br />

An analysis of an ice floe impact on a sluice gate will first be<br />

presented to illustrate the methods.<br />

The particular case of a "blocky" iceberg colliding with either a<br />

cylindrical or a conical gravity platform will then be analyzed.<br />

For typical iceberg characteristics and structure dimensions, it<br />

will be demonstrated that a gravity structure could be designed to<br />

withstand iceberg loadings. The effect of structure movements will<br />

also be considered.<br />

Canada<br />

Canada<br />

653


INTRODUCTION<br />

In recent engineering studies conducted by Acres it has been necessary to calculate<br />

the impact of large ice features on several different types of marine structures and<br />

a simple design approach was adopted. The approach is based on kinetic energy<br />

diSSipation, which is similar to that used by other researchers in the calculation<br />

of the impact of an ice floe hitting a pier (Michel, 1978), and the determination of<br />

iceberg scour (Chari, 1981).<br />

In these studies the following design assumptions have been used.<br />

- When an ice feature collides with a massive structure, the leading edge of the ice<br />

feature will be progressively crushed, and the force acting on the structure<br />

during impact will be that of a steady increase from zero to a certain maximum<br />

value, as a result of kinetic energy dissipation during the ice crushing period.<br />

- The shape of the ice feature has been arbitrarily simplified for preliminary<br />

calculations.<br />

- The deformation of the structure has been neglected, since it is normally an<br />

insignificant portion of overall displacement.<br />

- Ice feature dimensions and strengths are such that buckling and bending failures<br />

will not occur, and a <strong>com</strong>pression or crushing failure is assumed.<br />

- For large ice features such as iceberqs the volume of water in motion with the ice<br />

feature will have influence on impact energy. The virtual displacement of the ice<br />

feature is obtained by adding this volume to the ice feature displacement.<br />

The following analysis represents two different case studies I that of an individual<br />

ice floe in collision with a sluice gate, and an iceberg impact with an offshore<br />

gravity platform.<br />

ICE FLOE IMPACT AGAINST<br />

A SLUICE GATE<br />

When an ice sluice gate (Figure 1) is partially open the water level overtops it.<br />

At this moment an ice floe which moves with velocity V can strike the top of the<br />

gate and create an impact load F in addition to the hydrostatic load. The collision<br />

654


where k is assumed to have a value of 1.5. This coefficient is <strong>com</strong>monly used for<br />

calculation of the hydrodynamic mass of ships in berthing energy calculations<br />

(Canadian Coast Guard, 1977), and for nblocky· shapes, it is likely a reasonable<br />

estimate.<br />

The maximum value of penetration "m can be determined by equating the kinetic<br />

energy of the ice floe to the energy absorbed by the crushing of the icel<br />

wv 2 = 1. 5Wi v2<br />

2g "2"q<br />

Bfcr<br />

cosa<br />

"m<br />

JXdx o<br />

where Ek = kinetic energy of the ice floe<br />

Hence,<br />

Ec = energy dissipated in crushing.<br />

and the maximum impact force Fm is expressed as<br />

Representative values of all variables for a typical case study are given in<br />

Table 1.<br />

Table<br />

Impact on Sluice Gate<br />

Ice floe dimensions<br />

Ice floe displacement<br />

Ice floe velocity<br />

Ice crushing strength<br />

Sluice gate slope<br />

Maximum penetration<br />

Maximum impact force<br />

656<br />

B = 5 m, L = 15 m, D 1 m<br />

Wi = 67.5 tonnes<br />

V = 0.20 m/s<br />

fer - 1.0 MPa<br />

c:a 70·<br />

"m = 1.7 em<br />

Fm 24.4 tonnes<br />

(4)<br />

(5)<br />

(6)<br />

(7)


For a cylindrical platform, the maximum impact force Fm is determined as follows:<br />

It is assumed that the arc length (acb) is approximately equal to the chord length<br />

(ab), if the total penetration is small.<br />

contact area is (Figure 2)<br />

where D = iceberg depth<br />

R = platform radius<br />

L<br />

I BLOCKY I BERG<br />

ELEVATION<br />

PLAN<br />

Figure 2 - Iceberg Impact on Cylindrical Platform<br />

658<br />

For a penetration of distance x, the<br />

CYLINDRICAL<br />

PLATFORM<br />

(9)


The impact load is then defined as<br />

and the energy dissipated during crushing is<br />

Xm<br />

Ec = 2Df crJi( 2Rx - x 2 )1/2 dx (11 )<br />

o<br />

Equations (8) and (11) can be solved numerically, or an approximate expression for<br />

Xm can be found by dropping the x 2 term in (11) (if >


Again, for small penetrations, it is as'sumed that the area (abck) in Figure 3 is<br />

equal to the projected parabolic area (akc), so that<br />

AX = 2. (ac)(kf) =.! tanSx<br />

3 3<br />

and Rl is the platform radius at the point of impact, which is<br />

R _ (d - di)<br />

tanS<br />

where R radius of platform base<br />

base angle<br />

water depth<br />

depth of iceberg impact<br />

The parameters Fx ' E c ' xm and Fm are defined as<br />

.! tanSfcrx ( 2Rlx - x2 )1/2<br />

3 '<br />

(2R - x 2 ) 1/2dx<br />

I<br />

! tanSfcrXm ( 2Rlxm - Xro2 )1/2<br />

3<br />

Of particular application to possible production platforms for the Hibernia<br />

,<br />

discovery, the numerical values given in Table 2 are substituted. The platform<br />

dimensions are those for an approximate water depth of 80 m and are similar to those<br />

discussed bY Jarlan (1981). The' maximum credible' iceberg anticipated for this<br />

depth can have a displacement of approximately 1.2 x 10 7 tonnes. The above<br />

analysis indicates that for the cylindrical platform, the expected impact force is<br />

7.6 x 10 5 tonnes, and for a conical platform is 1.4 x 10 5 tonnes.<br />

660<br />

( 17)<br />

( 18)<br />

( 19)<br />

(20)<br />

(15)<br />

( 16)


Table 2<br />

Design Example, Iceberg Impact<br />

on Gravity Platform<br />

Iceberg displacement, Wi<br />

Iceberg depth, D<br />

Iceberg draft, di<br />

Water depth, d<br />

Iceberg velocity, V<br />

Ice strength, fcr<br />

Platform radius at base, R<br />

Platform base angle,<br />

Maximum penetration, xm<br />

Maximum impact force, Fm<br />

Platform displacement (est)<br />

Coefficient of friction<br />

Factor of safety versus sliding<br />

Cylindrical Platform<br />

1.2 x 107 tonnes<br />

75 m<br />

60 m<br />

80 m<br />

1.0 mls<br />

5.0 MPa<br />

60 m<br />

90·<br />

0.88 m<br />

7.6 x 105<br />

1.1 x 10 6<br />

0.70<br />

1.01<br />

tonnes<br />

tonnes<br />

Conical Platform<br />

1.2 x 107 tonnes<br />

75 m<br />

60 m<br />

80 m<br />

1.0 mls<br />

5.0 MPa<br />

100 m<br />

40·<br />

7.81 m<br />

1.4 x 10 5 tonnes<br />

1.1 x 10 6 tonnes<br />

0.70<br />

5.35<br />

Again simplifying the design process, the factor of safety against sliding is<br />

calculated for each case, as indicated in Table 2. This shows that the =nical<br />

platform, in particular, could be designed to resist iceberg impact forces. If the<br />

structure can be designed to ac<strong>com</strong>modate limited horizontal movement, a higher<br />

factor of safety can be assured.<br />

Naturally these calculations are preliminary only, since the nature of iceberg-<br />

structure interaction is still poorly understood. A more detailed study would<br />

involve a much more careful analysis of iceberg displacements, velocities and<br />

strengths. The nature of the contact face during impact is especially sensitive, as<br />

the above analysis demonstrates.<br />

CONCLUSIONS<br />

Using a simplified model for ice-structure interaction, it is possible to develop<br />

impact forces on offshore and marine structures. The analysis is based on individ­<br />

ual ice features colliding with structures, where the crushing of ice dissipates the<br />

kinetic energy of the ice feature. The analysis indicates as an example, that large<br />

gravity platforms for the Hibernia field could possibly be designed to absorb ice-<br />

berg impact. However, a more careful evaluation of the design variables must be<br />

performed to achieve more confidence in the results.<br />

661


ACIQIOWLEDGMEN'l'S<br />

The authors wish to thank Mr. David B. Sampson, General Manager of Acres-Santa Fe,<br />

for his support of this publication.<br />

REFERENCES<br />

Canadian Coast Guard (1977)<br />

"Termpol Code-Code of Re<strong>com</strong>mended Standards for Prevention of Pollution in Marine<br />

Terminal Systems"<br />

Transport Canada, 1977<br />

Chari, T. R. (1979)<br />

"Geotechnical Aspects of Iceberg Scours on Ocean Floors"<br />

Canadian Geotechnical Journal, Vol 16, No.2, pp 379-390.<br />

Jarlan, G.E. and Lehalleur, J.D.<br />

"A Prestressed Concrete Fixed Drilling and Production Platform for the Hibernia Oil<br />

Field Development"<br />

Symposium on Production and Transportation Systems for the Hibernia Discovery,<br />

St. John's, February, 1981<br />

Lewis, J. (1981)<br />

·Icebergs on the Grand Banks: Oil and Gas Considerations"<br />

World Oil, January 1981<br />

Michel, B. (1978)<br />

"Ice Mechanics·<br />

Les presses de l'Universite Laval, Quebec, 1978<br />

662


M. Rojansky<br />

B. C. Gerwick<br />

Abstract<br />

Failure Modes and Forces of Pressure Ridges<br />

acting on Cylindrical Towers<br />

University of California, Berkeley USA<br />

This study is directed at the potential failure modes and sequences of failure<br />

of a typical sea ice pressure ridge impinging against the vertical cylindrical<br />

shaft of an offshore gravity platform. From these analyses, the maxima forces<br />

acting on the structure can be bounded.<br />

It was found that in the case of short ridges, the sheet behind the ridge will<br />

fail in crushing and create a rubble field. For longer ridges, the failure will<br />

be due to flexure. The critical length is a function of the geometry and kinematics<br />

of each given situation.<br />

Since actual pressure ridges will often have both consolidated and unconsolidated<br />

zones, a transformed section has been adopted.<br />

The indentation problem is addressed analytically, and the indentation factor<br />

is shown to vary non-linearly with the aspect ratio, being <strong>com</strong>Pl'-rable to previously<br />

published empirical data for both large and small aspec1l ratios but lying<br />

below them for intermediate ratios.<br />

A ridge which is incorporated within an isolated ice floe may exert a significant<br />

impact force when it impinges on a structure. The analysis for impact<br />

considers a series of increments for each of which the variations in apparent<br />

crushing strength due to changing strain rates and aspect ratios are considered.<br />

A numerical example is presented in order to demonstrate the principle<br />

that the "minimum required energy to failure concept" should be adopted in<br />

order to determine the mode or modes of failure and the maxima forces<br />

imposed on the structure.<br />

Introduction<br />

Continued studies of the potential interaction between sea ice and structures<br />

in Arctic and sub-Arctic regions have demonstrated that pressure ridges are a<br />

dominant feature in determining the maxima forces to be resisted. These ridges<br />

consist of both consolidated and unconsolidated ice and move with different<br />

velocities depending on the amount of open water.<br />

663


Although many previously published papers have re<strong>com</strong>mended that a vertical<br />

cylindrical shaft was unsuitable for an environment which includes multi-year<br />

pressure ridges, due to the alledgedly high forces developed by the ice in the<br />

crushing mode; however, the vertical tower has a number of advantages for<br />

specific sites, including the minimization of lateral and vertical wave forces,<br />

low wave run-up, relative freedom from problems occasioned by adfreeze and<br />

ice over-ride. Therefore it seemed worthwhile to reexamine the modes of<br />

failure and determine an envelope or upper bound of the forces actually developed.<br />

Failure Seguence<br />

Previously published reports of structures in Cook Inlet and the Baltic Sea,<br />

as well as model tests show that when a pressure ridge impinges against a vertical<br />

indentor, (Fig. I) the principal failure of the ridge does not occur in crushing of<br />

the contact zone between the ice and structure, but most often occurs in flexural<br />

cracking opposite or even at some distance laterally from the structure. Such<br />

cracking may occur repetitively and may result in a pile up of broken ice, ie. a<br />

rubble field behind the ridge. A sequence of failure modes is observed, of which<br />

<strong>com</strong>plete failure in crushing is seldom if ever dominant. This contrasts with the<br />

failure of a uniform ice sheet against such a vertical shaft, which is characterized<br />

by crushing and shear.<br />

This phenomenon of multi-modal failure is an example of the "minimum<br />

energy to failure concept" and is the underlying assumption behind the modal<br />

analyses presented in this paper. The ridge is assumed to fail first at a mode<br />

and at a location which requires the least amount of energy to failure. Subsequent<br />

failures will then similarly be determined, based on the structural and<br />

strength characteristics of the new ice formation.<br />

One way to identify a possible failure sequence is to assume various possible<br />

failure mechanisms and search for the critical one. The resulting failure mode<br />

and force will then be a function of the mechanical properties of the materials<br />

involved as well as the geometry and kinematics of the event. Particular attention<br />

is directed in this study to failures in crushing and in flexure.<br />

Mechanical Properties of Ice<br />

The mechanical properties of ice are obviously critical to any study of<br />

failure mechanisms and imposed forces. Although ice varies widely, there<br />

has fortunately been a great deal of research and testing of mechanical properties<br />

in recent years, much of which has been published. For the purpose of this<br />

paper, relevant information from references 3, 9 and 10 will be used.<br />

The most important aspects of mechanical properties to this study are:<br />

a. The crushing strength varies with the rate of loading (strain rate)<br />

such that three behavior zones, ductile, transition, and brittle are<br />

identified. (Fig. 2)<br />

b. Sea ice is approximately 5 times as strong in <strong>com</strong>pression as in<br />

flexure or tention. (Fig. 3)<br />

To simplify the analysis, the following assumptions have been made:<br />

664<br />

3.. Although sea ice is anistropic and non-homogeneous, the analytical<br />

model can be based on homogeneous and isotropic behavior because<br />

the failure against a vertical cylinder is essentially planar, and the<br />

structure of a multi-year ridge consists essentially of crystals<br />

oriented as vertical colwnns.


Under the assumptions described earlier the maximum positive moment will<br />

occur at the point of interaction and will be equal to:<br />

... p COShAL -COS)'L<br />

M : 4>.)( SlnhAL +SII'l).L eq.3<br />

The location of the maximum negative moment will vary, however<br />

for short ridges its magnitude may approach the magnitude of the maximum<br />

positive moment. For these cases the maximum negative moment will occur<br />

at the end of the ridge and will be given by:<br />

- P sinh L. SI " L<br />

M =-X)( 51'1 AL+SII'1). eq.4<br />

The failure force P can be found by equating the maximum externally<br />

applied moment to the maximum internal moment capacity.<br />

Ml1llx= -¥-<br />

where: 17£ The flexural strength of ice<br />

C The distance to the extreme "fiber"<br />

Once the initial crack has formed, the ridge separates into two shorter<br />

ridges which are assumed to be simply supported at the center (Fig. 4 b).<br />

The new maximum moment is most likely to occur at the fixed end and it is<br />

equal to:<br />

The resulting interaction force is found as described earlier, using<br />

eq.5. A <strong>com</strong>parison between the two interaction forces (P versus P')<br />

shows that for all practical cases, P' is smaller than P. Hence, for the<br />

stipulated flexural failure the maximum interaction force is associated<br />

with the formation of the initial crack.<br />

Analysis for Failure in Crushing<br />

Crushing failure can be defined as a failure in shear due to excessive<br />

<strong>com</strong>pression. To examine this failure mechanism we will evaluate the<br />

case of ice impinging against a flat indentor. This is a two dimensional<br />

idealization of a three dimensional problem. However, the same logic<br />

could be applied in the latter case. Several studies have been conducted<br />

in order to find relevant expressions which make it possible to predict<br />

the magnitude of the ice-structure interaction force when crushing failure<br />

takes place (Refs. 1,4,6,8). The various proposed relationships are<br />

similar to each other: thus here we use an expression from Ref. 8. This<br />

expression for the interaction force is as follows:<br />

F = fcX1"v>


SUMMARY<br />

The previous discussion offers a rational approach towards the analysis of<br />

the ice-structure interaction problem. In analyzing the interaction forces, the<br />

minimum required energy to failure concept can be beneficially adopted rather<br />

than considering the maximum force due to a single failure mode. Particular<br />

attention should be given to the development of a rational failure sequence.<br />

Acknowledgement<br />

The study which formed a basis for this paper was performed under a grant by<br />

Mobil Research and Development Corp., whose support is gratefully acknowledged.<br />

REFERENCES<br />

1. Acres/Santa Fe-Pomeroy - Feasibility study offshore drilling in the<br />

Beaufort Sea, APOA 1971<br />

2. Bercha, F. G., Stenning, D. M. - Arctic offshore deepwater ice-structure<br />

interactions, OTC 3632<br />

3. Blenkarn, K. A. - Measurements and analysis of ice forces on Cook Inlet<br />

structures, OTC 1261<br />

4. Croasdale, K. R. - Ice forces on fixed rigid structures - Calgary,<br />

Canada, 1978<br />

5. Hetenyi, M. - Beams on elastic foundation, University of Michigan<br />

Press, 1971.<br />

6. Korzhavin, K. N. - Action of ice on engineering structures, Siberian<br />

Dept. of the Head of Science, Novosibirsk, USSR 1962.<br />

7. Michel, B. - Ice pressure on engineering structures, CREEL, U. S.<br />

Army, Hanover, NH 1970.<br />

8. Ralston, T. D. - Sea ice loads - Technical Seminar on Alaskan Beaufort<br />

Sea Gravel Island Design, EXXON, Houston, TX 1979<br />

9. Vaudrey, K. D. - Study of related properties of floating sea ice sheets<br />

and summary of elastic and viscoelastic analysis - U. S. Navy,<br />

Civil Engineering Laboratory, Port Hueneme, CA 1977<br />

10. Wang, Y. S. - Sea ice properties - Technical Seminar on Alaskan<br />

Beaufort Sea Gravel Island Design - EXXON, Houston, TX 1979.<br />

670


of determining the total depth of disruption of the soil below the score exists and<br />

this will be the topic of future research. Multi-year ice, ice islands or ice<br />

island fragments generally cause a single, or a few, smooth, wide furrows, while<br />

first year keels cause multiple rake-like furrows in the sea floor.<br />

In the Canadian Beaufort Sea, a relatively stable land fast ice zone forms early in<br />

the winter and remains with little movement for most of the winter; little scoring<br />

would be expected in this zone during the winter season. Bordering this zone close<br />

to the 20 m water depth is the shear zone (Croasdale and Marcellus 1977) which<br />

separates the highly mobile pack ice further offshore and the landfast ice; significant<br />

scoring occurs each year in the shear zone. In deep water few scores occur, as<br />

only extreme keels are able to reach to the seabed and beyond 47 m it is believed<br />

that no current scoring is occurring, as a 47 m keel is the maximum keel that has<br />

been observed.<br />

Once a score has been formed, dependant on its cross-section it may, immediately, be<br />

partially filled-in by the scored sediments as illustrated in Figure 1. Other in­<br />

filling processes (shown in Figure 1) such as preferential infilling (Barnes and<br />

Reimnitz 1979) and superimposition (Lewis 1977a) could cause the disappearance of<br />

scores which is much greater than expected by infilling at the average sedimentation<br />

rate. Sedimentation and score infilling are qui te distinct processes. Uniform<br />

sedimentation by river sediments or sediments reworked by the scoring process results<br />

in minimal infilling of the scores (Figure lB). Immediate infilling, superimposition<br />

and preferential infilling cause the most significant infilling of scores (see<br />

Figure 1).<br />

Immediate infilling is a function of the characteristics of the soils scored. The<br />

probability of superimposition infilling is a function of the frequency of scoring<br />

in a given area. Preferential infilling however may be the result of a number of<br />

physical environmental occurrences, such as normal and tidal currents, wave induced<br />

currents, ice keel induced currents or turbidity currents (which could be caused by<br />

earthquake activity). Thus in shallow waters, off the Mackenzie Delta score infill­<br />

ing will occur almost continuously, whereas in deep water, infilling will be episo­<br />

dic, occurring only during extreme environmental conditions; in very deep water<br />

probably no wave induced infilling occurs. Also scores behave like sediments traps<br />

on the seabed, whereby the seabed roughness dictates the quantity of sediment trap­<br />

ped in a specific area during a certain interval of time.<br />

Thus one would expect to observe very few scores on the seabed near shore due to<br />

both a low scoring rate and a high fill-in rate, a zone of maximum scoring in 15 to<br />

35 m water depths and numerous scores on the seabed in water depths beyond 40 m dne<br />

to a low fill-in rate.<br />

676


AVOIDANCE OF SCORING<br />

It is possible to avoid having to trench the pipeline by directing its route through<br />

areas that are not scored due to the local topography, i.e. route pipelines down<br />

natural depressions in the seafloor.<br />

There are several extinct river valleys or channels in the Beaufort Sea. It was<br />

thus felt that it may be possible to run a pipeline along these low points in the<br />

seafloor. A study (Marcellus 1980) indicated that it is not economical unless the<br />

valley is close to the required route. Alternative techniques of protecting pipe­<br />

lines in ice infested waters are discussed by Marcellus and Palmer (1979).<br />

METHODS OF DETERMINING PIPELINE BURIAL DEPTHS<br />

Burial Below The Saturated Scored Zone (SSZ)<br />

Based on the discussion in the previous chapter a pipeline set below the SSZ (in<br />

areas where seabed erosion is not occurring) would be below the deepest scour that<br />

has occurred since the transgression period for that particular water depth.<br />

In water depths where there is 10,000 years of sediment and scoring has occurred<br />

during the entire time period, the bottom of the SSZ would have been formed 10,000<br />

years ago and current scores may no longer reach to this depth. In shallow water,<br />

assuming no erosion, the sediments have been deposited for about 3,000 years, so the<br />

depth of the SSZ indicates the deepest scores that have occurred during this period<br />

of time. Thus in shallow and medium water depths the bottom of the SSZ is probably<br />

a reasonable TOP depth but in deep water it is probably unreasonably conservative.<br />

Burial Below The Deepest Scores<br />

Over the past few years, large quantities of data have been obtained from echo<br />

sounding and side scan sonar traces of scores of the seafloor. Echo sounding data<br />

gives the depths of scores and the deepest score along a pipeline route can be<br />

obtained from these data. Does the depth of the deepest score indicate a safe TOP<br />

depth? Figure 3 shows the deepest scores from APOA 32; more recent data are present­<br />

ed in APOA 122, but these are confidential.<br />

In extremely shallow water scores are rapidly obliterated due to the high level of<br />

hydrodynamic activity in this water depth. In deep water the observed scores are<br />

the result of scoring which occurred when the water level was some tens of metres<br />

below its current level and the resulting burial depth would then be extremely<br />

conservative. In the mid-range water depths where the scoring rate is high (around<br />

the shear zone) the deepest score which has occurred in the past 50 years may have<br />

678


een significantly infilled. Thus in these water depths burial below the deepest<br />

score is probably inadequate.<br />

Therefore it is felt that the maximum score depth cannot be used to select a safe<br />

TOP depth at this time. This method is useful for <strong>com</strong>parative purposes.<br />

Score Dating<br />

If scores could be dated directly, then it would be possible to calculate the<br />

return periods for scores of various depths. In this method, it is necessary to<br />

identify the original score surface and date the sediment immediately above this<br />

surface. In medium water depths where scoring is very active and scores are filled<br />

by the disturbed sediments from the SSZ, it is not expected that dating will work.<br />

In deep water where scoring is very infrequent and scores are probably filled by<br />

river sediment, dating may be possible. A score dating project by Dome in 1975<br />

(Dome, 1975) in intermediate water depth was unsuccessful, presumably for the reasons<br />

stated above.<br />

Repetitive Mapping<br />

In this method a given area on the seafloor is surveyed by Side Scan Sonar every<br />

year or every few years, and the number of new scores which have occurred during the<br />

interval are counted. This provides the number of scores per year per kilometer for<br />

the specific area. LeWis (1977b) reports on data obtained between 1971 and 1974<br />

near Pullen Island in the Canadian Beaufort Sea. His findings for the 15-20 m water<br />

depth showed 10 new scores during three years at one site and 11 new scores during<br />

two years at the other site. Considering the areas surveyed, an average number of<br />

0.35 ::!: .12 scores per year per nautical mile, or 0.19 ::!: 0.07 scores/year/km was<br />

calculated. Knowing the score depth distribution and the number of scores/year/km,<br />

the return period for scoring can be estimated by the follOWing equation, (Weeks, et<br />

al,1980);<br />

d - {In (l/NTL»/{-k)<br />

where: d is the score depth predicted (m), L is the length of pipeline route (km),<br />

T is the return period desired (years), N is the average number of scores/year/km, k<br />

is a parameter equal to the reciprocal of the sample mean score depth for a given<br />

water depth (m- 1 ), In is the natural logorithm.<br />

Using an N of 0.19 scores/year/km, k - 2.5 (from Lewis 1977b), T = 100 years and L =<br />

78 km, we find that for the water depth range 15-20 m the TOP depth for a pipeline<br />

should be 2.9 m. This is plotted in Figure 2.<br />

679


The main problem with this method is the short time base of available data and the<br />

resulting relatively poor score statistics. For a given area the yearly variations<br />

in the local ice features scoring the seabed affect the number of new scores seen<br />

during that period. In this respect an averaging of data over a long period should<br />

give a better representation of the scoring rate for a given area. Secondly this<br />

method uses the observed score depth distribution on the seabed with no correction<br />

for infilling of the new s'cores. Therefore based on the previous discussion this<br />

method may tend to under or over predict the required TOP depth for a pipeline for a<br />

given return period. We view these data as extremely useful since they are the only<br />

data which use direct measurements of current scoring to obtain TOP depths for<br />

specific return periods, but better statistics are required.<br />

Scoring Equilibrium Analysis<br />

The basis of this analysis (Lewis 1977b) is that the number of scores in an area is<br />

in steady state and thus the number of new impacts in the area equals the number of<br />

scores infilled plus the number of impacts superimposed on existing scores. For the<br />

infilling rate Lewis used the average "sedimentation" rate based on the total mud<br />

thickness at a given site divided by the time available for deposition of mud in<br />

that area. However, considering the previous discussions this is clearly not cor­<br />

rect.<br />

Lewis' (1977b) basic equation was used to obtain the following equation:<br />

where:<br />

d (In(V/(un o + (V pno)/k»/-k<br />

d is the predicted score depth for a return period of T years<br />

V = l/LT<br />

L is the total length of pipeline (km)<br />

u is the sedimentation rate (mm/year)<br />

no is the number of scores/km in the area of interest about to be<br />

filled in<br />

This equation was used to obtain the line shown in Figure 2. The data used in the<br />

analysis was Lewis' (1977b) data for a pipeline route from the Kopanoar Wellsite to<br />

shore which incidently goes through the Pullen sites surveyed by repetitive mapping<br />

(discussed above). It is seen that for the 15-20 m water depth range Lewis' analysis<br />

indicates a TOP depth which is less than the point for the repetitive mapping method,<br />

which is based on an observed scoring rate in the indicated water depth range. This<br />

would then indicate that the infilling rate used by Lewis is lower than the actual<br />

infilling rate occurring at the Pullen site. By fitting Lewis' method to the ob­<br />

served point from repetitive mapping a new infilling rate can be determined which is<br />

680


2.8 times the sedimentation rate used by Lewis. If the discussion in the chapter on<br />

origin and disappearance of seabed ice scores is correct, we would expect that the<br />

scoring equilibrium analysis as documented by Lewis would overpredict the amount of<br />

scoring in deep water (where fill-in rate is less than the sedimentation rate) and<br />

underpredict the amount of scoring in shallow water (where fill-in rate is far<br />

greater than sedimentation rate), which is indeed noted in <strong>com</strong>parison with the other<br />

method presented.<br />

If, in line with our previous discussion, a high fill-in rate is assumed in shallow<br />

water (Barnes and Reimnitz (1979) reported infilling of .6 m in one score in 13 m<br />

of water in one year), and very little fill-in in deep water, Lewis' method can be<br />

made to better fit the observations (Marcellus 1981). However, much more information<br />

is needed on score infilling rates before these results can be used to select<br />

safe TOP depths.<br />

Ice Keel/Score Statistics Method<br />

This method uses the measured keel depth distribution passing over a point on the<br />

sea floor (obtained by upward looking profiler, submarine profiler, laser profiler)<br />

and the measured scour depth distribution, observed on the seafloor.<br />

The number of keels greater than depth D metres passing a point on the seafloor/year<br />

can be represented by:<br />

-D/Do<br />

N()D} = Noe (c.f. Wadhams 1975)<br />

Where No is the total number of keels/year and Do is a constant.<br />

The number of keels pasing over a 1 km long pipeline is f times the above, where:<br />

f 1000/ (w/2)<br />

where w is the width or length of the keel (m). W can vary from the length of a<br />

keel (approximately 2000 m) to width of keel (D/tan 35°), where D is the ice keel<br />

depth; here an average value for f is used.<br />

The distribution of scores of depth d on the sea floor is given by:<br />

e- kd<br />

(Lewis 1977b)<br />

Thus the probability of a score d metres deep occurring in a water depth of D is<br />

given by:<br />

681


-0/00 -kd<br />

p(d) a Noe fe scores/year/km of depth d<br />

This method uses current ice feature data, features which must be causing the curr­<br />

ently observed scoring. It allows for extreme events provided they are from the<br />

observed keel distribution (one can use submarine data for the area North of Dome's<br />

well sites in the permanent pack ice and also up along the Banks Island coast which<br />

show extreme features); here the keel distribution is from Wadhams (1975) laser<br />

data to demonstrate the method as it is public. The model assumes that the distribu­<br />

tion of score depths is constant with time. This is true if all scores fill-in<br />

uniformly with time. It is conservative if the shallow scores fill in faster than<br />

the deep scores and non-conservative if the deep scores fill in faster; in fact, the<br />

assumption is probably conservative. The factor f is not known, but can be bracket­<br />

ed. The method also allows for the different soil types and local bathymetry, as it<br />

uses measured score depth distributions.<br />

Thus the TOP depth for a pipeline of length L and return period T is given by:<br />

d = In (LT N f e-D/Do)/k<br />

o<br />

Assuming a pipeline of length 78 km, a value of No of 2300 keels/year derived from<br />

Wadhams (1975), Do = 2.5 and Lewis' (1977b), k for different water depths, the TOP<br />

depths as indicated in Figure 2 were obtained for 100 year return period.<br />

Here we have utilized pack ice ridge statistics to cover the entire route from shore<br />

to Kopanoar. This is acceptable at and beyond the shear zone; however, within the<br />

landfast ice, which is relatively stationary throughout the winter, these statis­<br />

tics over estimate the occurrence of scoring and hence TOP depths.<br />

Ice Keel-Soil Interaction<br />

Theoretical studies have been conducted (Kivisild et aI, 1981) to calculate the<br />

forces involved in scoring, the available environmental forces and the integrity of<br />

ice keels. The study shows that a narrow, deep multi-year keel being pushed by pack<br />

ice under extreme storm conditions could cause very deep scores in soft seabed<br />

soils.<br />

The difficulty of using the method is correctly quantifying the driving force on the<br />

ice keel which to date is the topic of extensive research, the probability of this<br />

force occurring, and the probability of the ice keel running aground. Estimating<br />

score return periods using this method would be difficult at this time.<br />

683


TOP Depth Optimization<br />

Optimum pipeline TOP depths should be based on minimizing the costs over the life of<br />

the pipeline. The total cost of the pipeline is given by the sum of installation<br />

and trenching costs plus the cost of a disruption times the probability of a disruption.<br />

Thus:<br />

Cost C = I + A(d + 0 )2 L + B p(d) LT<br />

where: I is the cost of installing the pipeline of length L, A(d + 0 )2 is the cost<br />

of trenching the pipeline, 0 m in diameter, d metres (TOP) into the seabed, B is the<br />

cost of a disruption and p(d) is the probability of a disruption.<br />

To minimize the total cost of the pipeline we set:<br />

dC/dd = 0 2A (d + 0) L + B P (d) LT/dd<br />

Using the expression for p(d) from the ice keel/score statistics method above we<br />

get:<br />

2A (d + 0) B k f N o<br />

Where T is the lifetime of the pipeline in years.<br />

-0/00 -kd<br />

Te e<br />

The results of evaluating this expression for A = $20,000 and B = $1.5 billion for<br />

a pipeline route from Kopanoar to shore are presented in Figure 3.<br />

DISCUSSION<br />

This paper presents methods of calculating TOP depths for protecting subsea pipe­<br />

lines in ice infested waters.<br />

The SSZ and deepest score depth data are obtained by taking observations of the sea<br />

floor. However, these data must be considered with reservations. It is felt that<br />

setting the TOP just below the SSZ in shallow to moderate water depth (depths where<br />

there is current scoring) is reasonable; but in deep water where no (or very little)<br />

active scoring is taking place, the thickness of the SSZ is a relic of former times<br />

and setting the TOP below this depth is far too conservative. A similar argument<br />

can be made for the deepest score. In shallow and medium water depth scores are<br />

filled in rapidly and the "deepest" score may be significantly reduced by infilling.<br />

In deep water the deepest score may be relic and setting the TOP below this depth<br />

would be too conservative. These two methods are felt to be useful for <strong>com</strong>parative<br />

purposes only in deep water, but the SSZ-is a reasonable TOP depth in shallow water.<br />

684


The score equilibrium analysis as proposed by Lewis is invalid if the sedimentation<br />

rate is used for score fill-in. It predicts a deep TOP in deep water and probably<br />

inadequate TOP depths in shallow water. However, if the score fill-in rate is<br />

reduced to zero in deep water (i.e., uniform sedimentation only) then this method<br />

can be made to give more reasonable TOP depths, but these values cannot be con­<br />

sidered quantitative or useful for engineering design at present.<br />

The repetitive mosaic work is extremely useful as this gives the only actual available<br />

information on current scoring. However, the method must be conducted for<br />

many years to generate adequate statistics.<br />

Score dating cannot be used in medium water depths where scoring is frequent, but<br />

may be useful for dating the deep scores in water depths of more than 40 m. This<br />

could be invaluable to rationalize no burial in deep water as indicated by the ice<br />

keel/score statistics method discussed here.<br />

At present it is felt that the keel/score statistics method is the most reliable.<br />

Recent results worked out by Dome utilize about 7 years of data (confidential) for<br />

the keel statistics. Site specific score depth distributions were used thus incor­<br />

porating the effects of soil type and local bathymetry.<br />

It is interesting to note that the results in Figures 2 and 3 show some agreement<br />

between the methods disscused. This, coupled with an understanding of the physics of<br />

the particular methods, provides credibility for the TOP depths indicated. The big<br />

discrepancy is the indicated TOP depth in deep water between the SSZ, deepest score<br />

and score equilibrium methods and the ice keel/score statistics methods. However,<br />

it has been rationalized here that the first three methods mentioned will greatly<br />

overestimate the required TOP depth in deep water. Likewise, the ice keel/score<br />

statistics method will overestimate the TOP depth in shallow water due to the data<br />

used and the ice morphology in the Beaufort Sea. It is felt that the "SSZ" depth<br />

should be used in shallow water and the ice keel/score statistics method in water<br />

depths beyond 20 m.<br />

One possible problem with ice score statistics methods of predicting the TOP depth<br />

is that the score depth may not indicate the total depth of disturbance in the<br />

seabed; however, this is not a problem with the SSZ method. This problem will be<br />

the topic of future research.<br />

In summary we feel that significant advances have been made in score research over<br />

the past few years, and statistical methods are being developed which will allow for<br />

sound engineering and safe design.<br />

685


LIST OF REFERENCES<br />

Arctic Petroleum Operator's Association (APOA) 1119, "Analysis of Records Showing Sea<br />

Bottom Scouring", prepared for Gulf Oil Canada Ltd.<br />

APOA 1132, "Analyze 1971 Scour Records And Run 1972 Program", prepared for Gulf<br />

Canada Ltd.<br />

APOA 11122, "Investigation Of Sea-Bed Scouring In The Beaufort Sea Phase IU", pre­<br />

pared for Gulf Oil Canada Ltd., by Maclaren Atlantic Ltd., December 1977.<br />

APOA 11151, "Ice Scour Mosaic Study", prepared for Gulf Oil Canada Ltd., by J. Shearer,<br />

May 1979.<br />

APOA 11158, "Ice Scour Mosaic Study", prepared for Gulf Oil Canada Ltd., by J. Shearer,<br />

May 1980.<br />

Barnes, P. and Reimnitz, E., "Ice Gouge Obliteration And Sediment Redistribution<br />

Event' 1977-1978 Beaufort Sea, Alaska", Report 79-848, U.S. Dept. Of The Interior<br />

Geological Survey, Menlo Park, California 1979.<br />

Blasco, S., 1980, Private Communication.<br />

Croasdale, K.R. and Marcellus, R.W., "Ice And Wave Action On Artificial Islands In<br />

The Beaufort Sea", The Canadian Journal Of Civil Engineering, Vol. 5, pp 98-113,<br />

March 1978.<br />

Dome (1975), "The Dating Of Ice Scours In The Beaufort Sea", Kenting.<br />

Kivisild, H.R., Pilkington, G.R. and Iyer, S.H., "An Analytical Study Of Ice Scour<br />

On The Sea Bottom", ASHE Conference, Houston, January 1981.<br />

Kovacs, A. and Mellor, M., "Sea Ice Morphology And Ice As A Geologic Agent In the<br />

Southern Beaufort Sea". CRREL, 1974, AINA Symposium On The Beaufort Sea Coastal And<br />

Shelf Research, San Francisco.<br />

Lewis, C.F .M., "The Frequency And Magnitude Of Drift-Ice Groundings From Ice-Scour<br />

Tracks In The Canadian Beaufort Sea", POAC 1977a.<br />

Lewis, C.F.M., "Bottom Scour By Sea Ice In The Southern Beaufort Sea", Dept. Of<br />

Fisheries And The Environment, Beaufort Sea Technical Report 23 (draft), Beaufort<br />

Sea Project, Victoria, B.C., 1977b.<br />

686


Marcellus, R.W., and Palmer, A.C., "Shore Crossing Techniques For Offshore Pipelines<br />

In Arctic Regions", POAC '79, Trondheim Norway, August 13-17, 1979.<br />

Marcellus, R.W., "Preliminary Order-Df-Magnitude Cost Comparison Of Alternative<br />

Pipeline Routes For Similar Pipelay And Trenching Methods From The Kopanoar Well­<br />

site To Shore", Canada Marine Engineering Report, prepared for Dome Petroleum Limited,<br />

June 1980, (proprietary).<br />

Marcellus, R.W., "Ice Score Return Periods And Pipeline Trench Depths In The Canadian<br />

Beaufort Sea", Canada Marine Engineering Report, prepared for Dome Petroleum Limited,<br />

April 1981, (proprietary).<br />

O'Conner, M.J. and Associates Ltd., "Development Of A Proposed Model To Account For<br />

The Surficial Geology Of The Southern Beaufort Sea", March 30, 1980.<br />

Wadhams, P., "Sea Ice Morphology In The Beaufort Sea", Beaufort Sea Project, Techni­<br />

cal Report Number 36, December 1975.<br />

Weeks, W.F., Barnes, P. W" Rearic, D., and Reimnitz, E., "Statistical Aspects Of<br />

Ice Gouging On The Alaskan Shelf Of The Beaufort Sea", (draft), CRREL, New Hampshire,<br />

1980.<br />

ACKNOWLEDGEMENTS<br />

The authors are very grateful to Bruce Dunwoody for his contribution to the develop­<br />

ment of the optimization method discussed above. The authors would also like to<br />

acknowledge Dome Petroleum for supporting the work upon which this paper is based<br />

and Beth Weber for the patience she has shown in typing this paper.<br />

687


IoKlDEL TESTS OF<br />

SEA BOTTOM SCOURING<br />

R. Abde1nour, Vice President Arctec Canada Limited Canada<br />

D. Lapp, Project Engineer Arctec Canada Limited<br />

S. Haider, Senior Marine Geotechnical Engineer, Petro Canada<br />

S.B. Shinde, Senior Research Engineer Esso Resources Canada Limited<br />

B. Wright, Coordinator, Frontier Development, Gulf Canada<br />

Canada<br />

Canada<br />

Canada<br />

Canada<br />

ABSTRACT<br />

A series of model scale tests of ice sea bottom scouring were conducted as part<br />

of the Arctic Petroleum Operator' s Assoc ia t ion project to acquire experimenta 1 data on:<br />

°scouring resistance forces<br />

°pressure distribution in the soil<br />

°pressure distribution on the m0de1 front face<br />

°shape and characteristics of the scour profile.<br />

The test variables included two model scales of 1 :25 and 1 :50, two model shapes<br />

an inverted pyramid and a prism; three soil materials, sand, sandy silt and silty<br />

clay. Each of these parameters were tested at three cut depths in a levelled soil<br />

bottom with three towing velocities.<br />

1 • INTRODUCTION<br />

The objectives of the test program were:<br />

°To determine the resistance force required for an ice mass to scour under<br />

various conditions of shape of the ice mass, soil type, cut depth and<br />

towing velocity.<br />

°To determine the pressure distribution measured on the ice model front face<br />

as well as within the surrounding soil in both horizontal and vertical dimensions<br />

related to the model.<br />

°To observe the behaviour of the soil during the scouring procedure and determine<br />

the scour profile characteristics behind the ice mass relative to its<br />

shape and cut depth.<br />

°To correlate the experimental results obtained with available published work<br />

including model and full scale tests.<br />

688


The test program provided significant information that covered most of the<br />

above objectives.<br />

The ice scouring tests were conducted at the ARCTEC CANADA LIMITED hydraulic<br />

basin in Kanata, which is 18.3 m long, 6.1 m wide and has a depth of 1 meter. The<br />

tests consisted of towing two model scale ice feature shapes "keels" constructed<br />

from steel plates and instrumented with force and pressure sensors, through three<br />

sea bottom conditions.<br />

The program was divided into three phases, each phase relating to a different<br />

sea bottom soil type. Each phase included four to seven test days in which the<br />

following variables were investigated:<br />

°mode1 shape an inverted pyramid and rectangular prismatic shape<br />

°mode1 scale model scales of A = 50 and A = 25<br />

°cutting depth three cutting depths<br />

°towing velocity: three velocities.<br />

Table 1 is a summary of the <strong>com</strong>plete test program which describes the number of<br />

runs performed in each soil and the model shape used.<br />

TABLE 1 - SUMMARY OF TEST PROGRAM<br />

Vt1/bl SOil<br />

nST<br />

•<br />

NO. OF Rutl$<br />

/lK)OEl LJAn Of TESTING<br />

1979 PER TEST' PER SOil<br />

I Sand 1 Pyramid 25, 26 JUlie 9<br />

Safld Pyr'amld 28 June 9<br />

Sand 3 large Prismatic 3, 4, 5 July 6<br />

Sand 4 Small Pnsmatlc 6 July 9<br />

II Silt 5 Pyram1d 30 July 9<br />

Slit 6 Small Prismatic 21 August 9<br />

13. 18 Sept.<br />

SIlt 7 large Prismatic 16. 21 Sept. 1<br />

Silt 8 Pyramid 24 Sept. 4<br />

--<br />

III Clay 9 Pyramid 29 October 9<br />

Cldy 10 Pyramid 31 October 0'<br />

Clay 11 Snlll11 PrlSmatic 5 Noyen'ber 9<br />

Chy 12 Small Prismatic 7 Noventler 9<br />

Clay 13 large PriSAlatlc 9 Noventler 9<br />

Clay 14 large PrlSmatic 13 Novenber 9<br />

Clay 15 Pyramid·-Front<br />

face only 15 NO'i'eniler 9<br />

'I:ddl tests consisted of three mdentor cut depths and three velocities.<br />

Numilt!fS shown In lhis colwm are the ones Iiitlere data was obtained with<br />

tH.,02lJtahlt! accuri\Cv.<br />

'Nll1t: runs were conducted but results dre suspected due to improper preparation<br />

uf the c1dY.<br />

The <strong>com</strong>bination of these parameters would have resulted in 144 runs. However,<br />

some of the problems encountered reduced the test runs to 110. It was clearly not<br />

possible to investigate all possible "keel" shapes for economic reasons, and so the<br />

two shapes were chosen to represent extremes within which all other shape configurations<br />

were assumed to lie.<br />

33<br />

23<br />

54<br />

689


The indentor models were rigidly fixed to the carriage allowing only one degree<br />

of freedom. This implies that the ice mass being modelled is large relative to the<br />

scour force and that there is therefore no movement generated by the scour in other<br />

directions.<br />

Three discrete soil depths were chosen rather than a continuous slope for ease<br />

of soil preparation and interpretation of results. Three artificially prepared<br />

soils were used; the first was sand, the second was sandy silt (termed silt in<br />

the paper) and third was silty clay (termed clay in the paper). All soils<br />

were obtained from the Ottawa area.<br />

In order to fulfill the experiment objectives, the program was designed and<br />

executed to concentrate on two specific activities:<br />

1) The study of the magnitude and extent of horizontal and vertical stresses<br />

induced by a moving ice feature including:<br />

(A) Indentor Model - horizontal and vertical forces on model faces and total<br />

forces of the whole model; - <strong>com</strong>pressive pressure on front face at various<br />

locations.<br />

(B) Soil - lateral extent of total stresses beyond the scour track; - lateral<br />

extent of porewater pressure changes in the soil beyond the scour track;<br />

- the zone of pressure influence in the soil; - extent and magnitude of<br />

axial total soil pressure within the scouring dimensions.<br />

2) Visual documentation of the interaction between the indentor and the soil<br />

including:<br />

o failure and shearing patterns<br />

o surcharge behaviour and dispersal<br />

o eddy effects on soil<br />

o apparent and original scour depths<br />

o deposition of material along scour tracks.<br />

2. EXPERIMENT PREPARATIONS<br />

The preparation for the execution of the experiments consisted of the following:<br />

o design and construction of the ice scouring models (indentors)<br />

o design of the pressure transducers and piezometers<br />

o overall instrumentation<br />

o soils type selection and preparation procedures.<br />

2.1 Design Construction and Instrumentation of Ice Scouring Models<br />

Two model shapes were selected for the experiments, one an inverted pyramid and<br />

the other a rectangular prism. For the two scales, two models were required for the<br />

690


2.1.2 RectanguU7r FTismatic Models<br />

Two models were built to simulate a prismatic ice feature at 1:25 and 1 :50<br />

scales. The models differed in width and length by a factor of 2. The shape contrasts<br />

with the pyramid in that the sides are vertical. The basic construction of<br />

the prismatic model was similar to the pyramid except that there were three detached<br />

faces versus two. These three detached faces were connected to the rigid portion<br />

of the model through force blocks. The bottom face had a force block measuring in<br />

the vertical direction (2 axis). The force blocks for the side and front faces<br />

measured horizontal forces perpendicular to and parallel to the direction of travel­<br />

Y and X axis - respectively. A diagram of this arrangement is shown in Figure 3.<br />

Force blocks also measured the forces on the whole model in the X and Z directions.<br />

FIGURE 3 - SECTIONAL VIEW OF SMALL PRISMATIC MODEL<br />

JI- l PLANE<br />

L, CARRIAGE<br />

fORCE BLOCK ALLOCATION<br />

}-<br />

LOCATIONS OF PRESSURE TRANSDUCERS ON FRONT FACE<br />

OF SMALL PRISMATIC MODEL<br />

fRONT fACE<br />

693


Pressures on the front face of each model were measured in a similar manner to<br />

the pyramid model. However, the transducers were located in different positions as<br />

Figure 3 shows. The larger area of the front face permitted a transducer to be fitted<br />

at the corner of the model so that variation in pressure on the face could be<br />

measured in two different planes, if desired.<br />

2.2 Soil Description and Basin Preparation<br />

Three soils were used over the course of the test program. Each was selected<br />

on the basis of grain size analysis to simulate an ideal sand, silt and clay. Figure<br />

4 shows the grain size distribution of the three soils finally selected. The<br />

three materials were classified as fine sand, sandy silt and silty clay.<br />

Each soil was monitored and its properties logged. The following soil properties<br />

were determined over the course of the test program depending on soil type:<br />

°grain size - sieve and hydrometer (clay only)<br />

°bul k dens ity<br />

°water content<br />

°dry density<br />

°angle of internal friction (for sand and silt only)<br />

°Atteberg limits lfor sand and silt only)<br />

°shear strength by direct shear method, triaxial test, Swedish fall cone and<br />

cone penetrometer.<br />

FIGURE 4 - GRAIN SIZE DISTRIBUTIONS<br />

OF SOIL MATERIALS USED DURING PRESENT PROGRAM<br />

694<br />

,10 .. IV w.oeo.... 100 'III"IN"U 1 •• 1<br />

II I I I I II I I


Before testing <strong>com</strong>menced, the test basin was prepared so that some control<br />

over the properties of the soils could be achieved.<br />

This primarily consisted of constructing a filter for the soil to control its<br />

drainage characteristics. First, a bed of 25 mm stone was laid on the floor of the<br />

basin to a depth of approximately 0.17 meters. At each end of the basin a slot was<br />

left to act as a collecting area for the draining water. Water draining at either<br />

end of the basin was subsequently recirculated into the basin using a pump. Second,<br />

a mat of filter cloth was laid on top of the stone to screen out the soil but permit<br />

water to pass through into the stone.<br />

The soil depths in the basin were chosen to be approximately 30 cm deeper than<br />

the cut depth. The soil level depth was believed to be enough to dissipate any<br />

effect the floor of the basin might have had on the forces experienced by the models.<br />

The soil was laid in three stepped depths, each constituting a run, matching the<br />

planned scouring depths for each model.<br />

3. TEST EXECUTION<br />

3.1 Testing Procedures<br />

The testing program consisted of towing each of the three models in each of the<br />

three phases represented by the three soil types. In each phase a model was towed<br />

through three thicknesses of soil over three tracks, each of which represented a<br />

different towing velocity . Figure 5 shows the prismatic model being tested in the<br />

basin and Figure 6 shows the testing details in the basin.<br />

FIGURE 5 - THE LARGE PRISMATIC MODEL DURING TESTING<br />

695


FIGURE 7 - FLOW CHART OF TEST PROCEDURES<br />

No<br />

697


In addition to these lateral horizontal pressures perpendicular to the track,<br />

a single pressure cell was installed at the end of Run 3 within the cutting depth<br />

located on the centerline of the model along the axis of movement. The cell was<br />

placed so that it was just ahead of the point where the final test run was <strong>com</strong>pleted.<br />

After the installation of the model and the soil the instrumentation was<br />

<strong>com</strong>pleted, a check was made to see that all force blocks, pressure cells and<br />

piezometers were operational. Following this, the test was executed at the specified<br />

velocity.<br />

After the run was <strong>com</strong>pleted, the model was removed and reinstalled into the<br />

next test position. Soil instrumentation was removed and placed into the appropriate<br />

positions for the next run. This procedure was repeated until all three<br />

speeds had been run.<br />

The scours left by the three runs were then profiled and photographed.<br />

The soil was then reworked after all instrumentation in the soil was removed.<br />

The model just tested was disassembled and the force blocks and pressure cells were<br />

installed in the next model. The procedure just described was repeated for the new<br />

model with the reworked soil.<br />

After all the models had been tested, the soil was removed from the basin and<br />

a new soil was placed.<br />

3.2 Parameters Measured<br />

During and after test execution the following parameters were measured:<br />

°scour trench depths<br />

°resistance forces<br />

°pressure measurements<br />

°scour profiles.<br />

For scour trench depth, the original scour depth was recorded and <strong>com</strong>pared to<br />

the apparent past test depth. Resistance forces were measured in both horizontal<br />

and vertical directions including forces on the front face, bottom (for prismatic<br />

models) and one of the two sides. In addition, total along track horizontal and<br />

vertical forces on the model were recorded. The output of the resistance traces<br />

from the force blocks were recorded in analogue form on a 28 channel recorder and<br />

displayed on the oscillograph recorder at the same time. The peak average force<br />

for each run was reported. Figure 8 shows an example of the force versus time and<br />

penetrations.<br />

Pressure measurements were made using transducers to record pressures on the<br />

front face of each model at varying locations depending upon the model size and<br />

shape.<br />

698


The friction factor between the model and the soil was not varied during these<br />

experiments, therefore its effect on the measured force was not investigated.<br />

The relationship between the previous seven parameters is given by:<br />

where:<br />

f (q"<br />

F Inertial Force<br />

yL 3 Gravitational Force<br />

q,<br />

CiyL =<br />

Frictional Force<br />

Gravity Force<br />

Cohesion Force<br />

Gravity Force<br />

V2 External Force<br />

gL Gravity Force<br />

C<br />

yL<br />

Dimensionless Force<br />

Dimensionless Friction<br />

Dimensionless Cohesion<br />

Dimensionless Velocity<br />

This equation was used to obtain dimensional and nondimensional relationships<br />

for each group of results obtained for geometrically similar models tested in the<br />

same soil type. Therefore, the analysis of the data was done separately for each<br />

of the soil types and each of the two model geometries tested, pyramid and prismatic<br />

models.<br />

4.3 Conclusions<br />

Based on the results and their analysis the following have been obtained:<br />

°Empirical relationships describing the resistance force versus the soil<br />

properties and the speed of the indentor for all three soils tested. By<br />

<strong>com</strong>parison with other published work in this field, the results showed some<br />

agreement for the part of the work where similar studies have been undertaken.<br />

°The pressure data of the gauges installed at the indentor force, the pressure<br />

distribution and the average pressure at the front face were calculated and<br />

cOl'related to the total resistance measured at the same face.<br />

°The data of the soil Dressure gauges and the piezometers within the soil.<br />

Despite the small number of gauges used, the zero pressure distribution<br />

line, relative to the model location was abstracted.<br />

701


7. REFERENCES AND BIBLIOGRAPHY<br />

[1] Brooks, L.D., "Another Hypothesis about Iceberg Draft", POAC 79, Proceeding<br />

Volume 1, Page 24=152, Trondheim, Norway, August 1973.<br />

[2] Schuring, D.J. and Emori, R.I., "Soil Deforming Processes and Dimensional<br />

Analysis", Society of Automotive Engineers Report 897C, September 1964.<br />

[3] O'Callaghan, J.R., McCullun, P.J., "Soil Mechanics in Relation to Earth<br />

Moving Machinery", Proc. Inst. Mech. Engineers 1964-65.<br />

[4] Kovacs, A. and Mellors, M., "Sea Ice Morphology and Ice as a Geologic Agent<br />

in the Southern Beaufort Sea", in the Coast and Shelf of the Beaufort Sea,<br />

<strong>Proceedings</strong> of a Symposium on Beaufort Sea Coast and Shelf Research, December<br />

1974.<br />

[5] Chari, T. and Allen, J.H., "An Analytical Model and Laboratory Tests on Iceberg<br />

Sediment Interaction", IEEE International Conference on Engineering in the<br />

Ocean Environment, Volume I, August 1974.<br />

[6] Abdelnour, R. and Lapp, D., "Model Tests of Sea Bottom Ice Scouring", APOA<br />

Report 151, ARCTEC CANADA Final Report 356C-3, November 1980.<br />

705


Reidar Lien<br />

Geological Enqineer<br />

SEA BED FEATURES IN THE BLAAENGA AREA,<br />

WEDDELL SEA, ANTARCTICA<br />

Continental Shelf<br />

Institute<br />

Norway<br />

ABSTRACT<br />

Data for this contribution were gathered during two expeditions in the summer<br />

seasons 1976/77 and 197B/79, and consist of records with echo sounder and side-scan<br />

sonar. From these data we have constructed a tentative bathymetric map of the area,<br />

and the sea floor has been classified into four groups of sea bed features.<br />

Two of these are well known and widely described in the literature: plough marks<br />

from grounded icebergs, and conventional undisturbed sea bed. The other two, in<br />

spite of a <strong>com</strong>prehensive study, are not found in the literature. These features<br />

consist of a washboard pattern, and a hummocky, mosaiclike, sea bed pattern.<br />

The different features are described and shown on record sections. Further some<br />

record sections with special phenomena such as tracks of wobbling icebergs,<br />

arresting icebergs, multi-keeled icebergs etc. are shown. Finally the different<br />

patterns and phenomena will be discussed with reference to their process of<br />

formation.<br />

INTRODUCTION<br />

In 1976/77 and 1978/79 the Continental Shelf Institute participated in expeditions<br />

administrated by the Norwegian Polar Institute to the Weddell sea. Data for the<br />

following contribution were gathered during these expeditions. These data consist<br />

of registrations with echo sounder and side-scan sonar from the Blaaenga area on<br />

the Dronning Maud Land coast. The area with side-scan sonar profiles and tentative<br />

bathymetry is shown in Fig. 1. In 1977 this area was covered by the ice shelf, but<br />

in 1978 the shelf had calved and the area was free of ice.<br />

706


LEGEND:"-:-- Side-scan sonar profile lines with dots each 10 min.<br />

'-_ _ _ _ FIg. - Location of profile section shown in text<br />

uf<br />

300<br />

15°3(1<br />

Fig. 1 Approximate bathymetric map with location of side-scan sonograms and<br />

figure sections.<br />

SEA BED FEATURES<br />

In the area four groups of sea bed features have been registered :<br />

a) Washboard pattern with lateral ridges<br />

b) Unsystematic furrows<br />

c) Hummocky sea bed<br />

d) Undisturbed sea bed.<br />

Washboard pattern with lateral ridges<br />

An example of this feature is shown in Fig_ 2. The straight parallel stripes have<br />

different spacing and represent ridges on the sea floor. Most of the stripes seem<br />

quite persistent. The washboard pattern represent roughness on the sea floor, in<br />

the form of undulations. The spacing between the grooves, and the grooves themselves<br />

may be of different sizes. The grooves form different patterns because of<br />

their geometrical form, but normally they are parallel for long ranges.<br />

707


Fig. 4 Hummocky sea bed.<br />

DISTRIBUTION OF THE SEA BED FEATURES<br />

The different sea bed features are mapped along the profiles in Fig. 5. On the<br />

shallowest bank area the records are dominated by unsystematic furrows. The majority<br />

of these furrows disappear below 300-320 m, and the deepest registered was found at<br />

a depth of about 380 m below sea level.<br />

The hummocky sea bed feature is found mainly in the lee side slope of the bank in<br />

relation to the direction of movement of the shelf ice. It was also found in a<br />

trough near the southern border of the investigated area. The features are found in<br />

depths between about 290 and 350 m.<br />

The washboard pattern with lateral ridges has been recorded from the south- eastern<br />

slope of the bank. This pattern was recorded to a waterdepth of 400 m, but may<br />

extend deeper since there are no registrations from greater water depths. The upper<br />

limit of the feature is diffuse because there is a gradual transition to a sea<br />

floor dominated by the unsystematic furrows as the water be<strong>com</strong>es shallower. The<br />

furrows seem to be superimposed on the washbord pattern. This latter pattern has<br />

been registered up to about 280 m before the unsystematic furrows <strong>com</strong>pletely seem<br />

to extinguish them.<br />

Undisturbed even sea bed is registered only along one profile in the lower part of<br />

the western slope of the central bank of the area. The water depth for this pattern<br />

varies between 330 and 380 m along the profile.<br />

709


Another possibility is that no ridges were originally formed. The bathymetric map<br />

of the area is not very reliable, so if the mentioned ridge is sharp, it is possible<br />

that the icebergs have ploughed through it and transported the material in front of<br />

the iceberg all through the ridge.<br />

Fig. 6 Plough marks lacking the rims.<br />

However, this theory is more doubtful since we can follow large plough marks without<br />

lateral ridges for several hundred meters, and that would probably have given an<br />

excessive accumulation of material in the front of the icebergs.<br />

Some places the plough marks are flat and shallow <strong>com</strong>pared with its width (Fig. 7).<br />

The reason for this may be that possible earlier peaks and roughness have shedded<br />

during the grounding until the iceberg has got a shape that will be stable to<br />

further influences. The explanation may also be that the bottom of the iceberg<br />

represent the original bottom of the ice shelf.<br />

Forming of the washboard pattern with lateral ridges<br />

The washboard pattern is underlying the plough marks and must therefore be older. In<br />

the literature there is not any description of this pattern. However, similar<br />

patterns formed by single icebergs north of Alaska (1) have been described. Corresponding<br />

impressions are also recorded in this area (Fig. 7). Reimnitz et al. (1)<br />

explain the formation of this pattern by wobbling of the iceberg after its grounding.<br />

The impressions in Fig. 7 may also be explained in that way, and the spacing<br />

between the crossing ridges in the plough marks is due to the icebergs' stability<br />

and their drifting velocity. Further the scour may be directed by the topography<br />

like conventional plough marks.<br />

711


Fig. 7 Plough marks from wobbling icebergs.<br />

The washboard pattern with lateral ridges, however, seems to be quite unaffected by<br />

the topography. In Fig. 8 a sketch of the records from the three close spacing<br />

profiles in the south-eastern corner of the investigated area is shown (Fig. 1).<br />

This sketch is a line drawing from the profiles, and shows the main features from<br />

the records. The records nearly overlap. In this sketch we can, without too many<br />

assumptions, follow the parallel stripes for several hundred of meters without<br />

notable curvature. The elevation along the profiles is about 20 m.<br />

The direction of the parallel stripes seems to be the same on all profiles where<br />

they are registered, and the direction is at right angles to the barrier front. The<br />

pattern of crossing ridges and grooves between the parallel stripes (the washboard<br />

pattern) also seems to be persistent, and each groove is similar to the foregoing.<br />

The change in this pattern is gradual and often takes place during several hundred<br />

meters, or there is no change at all within one profile.<br />

The formation of this pattern is supposed to be similar to that described by<br />

Reimnitz et al. (1), but here the forces that cause the drift, and the drifting ice<br />

itself must have been of another magnitude.<br />

712


Fig. 8 Sketch of approximate mosaic.<br />

The ploughing ice might have been icebergs bounded by mighty sea ice or pressed<br />

together to a firmer body, but not closer than that the icebergs could have minor<br />

individual movements . The drifting forces may be wind and current acting on the sea<br />

ice and the icebergs, they may also be represented by the shelf ice acting on the<br />

ice mass .<br />

acting<br />

force<br />

1<br />

- II '<br />

A<br />

i \<br />

: \ ICEBERG<br />

! \<br />

.I I<br />

i J<br />

i I<br />

:--_J ________ _<br />

\<br />

equilibrium<br />

water surface<br />

-<br />

- ' -<br />

Fig. 9 Suggested formation of the washboard pattern with lateral ridges; A) the<br />

mechanism after grounding, B) resulting pattern.<br />

The pattern is thought to be formed by the pushing of the ice mass when the bottom<br />

of it is grounding on the sea floor (Fig. 9a). As the forces are tremendous, the<br />

ice mass is always pushed forward with its upper part. A point in the lower corner<br />

of the grounded iceberg will be the rotation center. As this motion goes on, the<br />

iceberg will be more and more unstable and the stress on the sea floor contact will<br />

8<br />

713


increase. When the action between the ice and seafloor is equal to the strength of<br />

the ice or sea bottom, or the friction between these, a break will occur at<br />

the contact, and the iceberg will move toward equilibrium. This jump may be<br />

initiated by tidal movements, but this is not supposed to be necessary. At each such<br />

jump there will be a ridge that delineate the front of the iceberg. In that way the<br />

grooves and ridges may be of different shapes, dependent on the shape of the iceberg's<br />

front. In Fig. 10 is shown a front of an arrested iceberg that would have<br />

formed an edged pattern, while the plough marks in Fig. 7 show smoother forms.<br />

Fig. 10 Plough mark from an iceberg with irregular lower front.<br />

The pattern may also have been formed by the back sides of the icebergs if the<br />

icebergs are floating with an inclination larger than the gradient of the sea floor.<br />

The gradient of the sea floor i s here about 1:150. Thi s supposed inclination may be<br />

due to ram s of the iceberg or other irregularities. The mechanism itself is thought<br />

to be similar to the foregoing description.<br />

When a front of such icebergs is moved along there will be icebergs of different<br />

sizes and stabilities in the ice mass, and the time when the break between sea<br />

floor and iceberg occurs, will be individual . In that way the space between<br />

the grooves may be different for each iceberg (Fig. 9b).<br />

The stripes or parallel ridges in the direction of motion of the icemass are thought<br />

to represent soil pressed up between the icebergs. If the position among the<br />

714


icebergs themselves and their form is steady, these ridges will be persistent. The<br />

small, less persistent, stripes may represent unevenesses of the icebergs' fronts<br />

or bottom like those of the ploughmark shown in Fig. 10. These stripes may be more<br />

local since the front or bottom of the iceberg that touch the sea floor may change<br />

by shedding, thawing, freezing or tilting of the iceberg as it is moved up the<br />

slope. These factors will also affect the geometry of the grooves.<br />

Origin of the hummocky sea bed features<br />

The hummocky sea bed features are generally found in lee side slopes. The features<br />

some places resemble the plough marks with crossing ridges which are shown in Fig.<br />

7, but closer investigations show no continuous ridges across the lines where the<br />

pattern to a certain degree shows orientation. The crossing lines are broken and<br />

reminds us of an unstructured mosaic feature. Reimnitz et al. (1). describe<br />

<strong>com</strong>parable imprints from the shelf outside Alaska. These are supposed formed in<br />

overconsolidated sediments at or near the surface by icebergs placing blocks of<br />

sediments in a random manner along its track.<br />

Since the feature concerned is generally recorded in slopes, there may also be an<br />

element of slide in the explanation. The sea floor might have been disturbed by<br />

icebergs, which have initiated slides or small sediment movements.<br />

The hummocky sea floor may also be explained by more or less sliding in areas with<br />

rich accumulation. These areas in the lee side slope of the bank are likely to be<br />

such areas. It is supposed that the shelf ice has been grounded on the bank before<br />

1979 and the areas lie in the distal side of a moraine accumulation that forms the<br />

top of the bank (Haisey, G.H, pers.<strong>com</strong>. 1980). The pattern then might have been<br />

formed during a rich supply of sediments by random distribution of the material and<br />

more or less local slides in the slopes.<br />

Undisturbed sea floor<br />

Along one profile the foregoing type of sea floor gradually changes to rather<br />

undisturbed soils. This profile is crossing the trough west of the central height<br />

of the bank, and the undisturbed sea floor is situated in water depths from about<br />

330 to 380 m. Why this area is apparently undisturbed by icebergs, may be because<br />

the water depths or locations have not been favourable. But this seems strange<br />

since iceberg plough marks are neighbouring this area to the west on corresponding<br />

depths. The transition between these features seems to be in the bottom of the<br />

trough of about 380 m water depth. Since plough marks have been formed in the western<br />

slope of the trough, plough marks have in all likelihood also been formed in the<br />

715


eastern slope of the trough. These marks might, however, have been blurred by fines<br />

from the depositing to the east, further up the slope. This sedimentation of fines<br />

is most likely concentrated from its place of release, downslope to the bottom of<br />

the trough, and gradually decrease upward the opposite slope.<br />

Age of the features<br />

The age of the plough marks is difficult to deduce, but from the registrations they<br />

seem recent, especially in the shallowest areas. However, the sedimentation of soils<br />

in the area generally seems to be sparse (Maisey, G.H. pers. <strong>com</strong>. 1980), so they may<br />

be quite old though they look fresh.<br />

Further, the ice shelf's natural depth in the area is measured by radio echo sounder<br />

and side-scan sonar (2). These measurements revealed a depth of the ice shelf of<br />

about 200 m below sea level. This may imply that few of the iceberg plough marks are<br />

made by local icebergs in recent time, since the shallowest height in the area is<br />

about 240 m. In the south-western corner of the area, where the water depths are<br />

less than 160 m, the ice shelf was grounded on the sea floor in 1979. Accordingly<br />

the plough marks, if they are recent, are presumably made by icebergs drifted from<br />

regions with thicker shelf ice, or remnants of icebergs which have got a greater<br />

draught than the ice shelf. If the plough marks are ancient, however, they may also<br />

be originated in times with lower sea levels.<br />

As to the washboard pattern with lateral ridges, this is clearly formed before the<br />

plough marks, since these gradually blot out the pattern in water depths from about<br />

300 m and shallower. Further this pattern is probably formed during lower sea levels,<br />

because the pattern is registered at water depths of at least 400 m referred to the<br />

sea level of today. And masses of icebergs with draughts of more than 400 mare<br />

rather improbable.<br />

LITERATURE<br />

(1) REIMNITZ, E., BARNES, P.W. and ALPHA, T.R., 1973: Bottom features and processes<br />

related to drifting ice on the Arctic shelf, Alaska. Department of the interior<br />

United States Geological Survey.<br />

(2) FOSSUM, B.A. and KLEPSVIK, J.O.: Studies of icebergs and iceshelf using sidescanning<br />

sonar. In prep.<br />

716


is required for tests inside bore holes. In site-specific foundation studies, static<br />

penetrstion tests appear to be extremely relevant and useful.<br />

With a view to develop a quick and economical way of testing surficial seafloor<br />

soils, a free fall penetrometer with 4S cm 2 nominal cross section, a 60· cone and a<br />

62S cm 2 friction sleeve was developed at Memorial University. A description of the<br />

penetrometer and results from its sea trials are reported by Chari et al (1978, 1979)<br />

As a part of the ongoing research on penetrometers, quasi-static and free fall tests<br />

were conducted in the laboratory to facilitate an interpretation of the data from the<br />

marine penetrometer. This paper relates to the quasi-static tests with the standard<br />

10 cm 2 "Fugro" type penetrometer and the 4S cm 2 Memorial University Penetrometer.<br />

A study of the two penetrometers in the free fall mode has been reported by Chaudhuri<br />

(1979).<br />

EXPERIMENTAL ASSEMBLY<br />

The layout of the experimental facility is shown in Fig. 1. A detailed descrip­<br />

tion of the assembly is given by Abdel-Gawad (1979). The penetrometer was activated<br />

718<br />

FIG 1: THE EXPERIMENTAL FACILITY


through a hydraulic system and was set to move at 20 _/sec in accordance with ASTK<br />

re<strong>com</strong>mendations. The output from the penetrometer was continuously recorded on an<br />

analog recorder. The soil target was prepared in a steel bin which was a 1 m cube<br />

open at the top. Silica sand and modelling clay were used as the representative sam­<br />

ples of cohesionless and cohesive soils. Table 1 gives the physical properties of<br />

the soils used.<br />

Comparison was made of the experimental results and with the different theoret­<br />

ical analyses available in the literature. Fig. 2 shows a summary of the different<br />

soil failure mechanisms suggested in the literature. The theories of Meyerhof<br />

(1961), Nawatzki and Karafiath (1978) and Durgunoglu and Mitchell (1975) were used in<br />

this analysis.<br />

De Beer (1945)<br />

Meyerhof (1951,1961)<br />

TEST RESULTS ON SAND<br />

Terzoghi (1943)<br />

Nowotzki and Korofiath<br />

(1972, 1978)<br />

Q<br />

FIG. 2: DIFFERENT MODES OF SOIL FAILURE<br />

Q<br />

Biarez et 01 (1961)<br />

Hu (1965)<br />

Durgunoglu and Mitchell<br />

(1973. 1975)<br />

Typical experimental results for dry silica-70 sand are presented in Fig. 3.<br />

Similar results were obtained for the different types of soil. Theoretical values<br />

using the peak friction angle from triaxial tests are also presented on the same<br />

figure. The analysis of these tests leads to the following conclusion:<br />

719


Theoretical values of Nowatzki and Karafiath are found to be consistently lower<br />

than the experimental values. The difference increases with increasing depths of<br />

penetration. This theory is the most conservative of all.<br />

Values obtained from the theory of Durgunoglu and Mitchell are the nearest to the<br />

the experimental values. This is true for the various soil types and penetrometer<br />

variables. These values are intermediate between the values obtained by the other two<br />

theoretical methods. This can be explained with reference to the failure mechanism<br />

shown in Fig. 2. It can be seen from the figure that the failure mode suggested by<br />

Durgunoglu and Mitchell is an intermediate case between Meyerhof' s failure shape and<br />

that of Nowatzki and Karafiath.<br />

TESTS ON CLAY<br />

To choose an appropriate theoretical analysis for cohesive soils, results of the<br />

penetration tests performed on the clay target were <strong>com</strong>pared with the theoretical<br />

values of Meyerhof (1961) and Durgunoglu and Mitchell (1975). The numerical technique<br />

suggested by Nowatzki and Karafiath (1978) is not applicable for cohesive soils.<br />

Typical results of penetration tests in clay are presented in Fig. 4.<br />

Theoretical values obtained by Meyerhof' s theory and that of Durgunoglu and Mitchell<br />

are also shown on this figure.<br />

It may be seen from the data presented that Meyerhof' s theoretical values are<br />

greater than the experimental values for relative depths greater than 4 and less than<br />

the experimental values for D/B less than 4. A similar phenomenon was observed and<br />

explained previously in the case of cohesionless soils. The values of Durgunoglu and<br />

Mitchell were found to be in better agreement with the experimental values for stiff<br />

clay and medium stiff clay. But for soft clay the agreement is not good. The theo­<br />

retical values was found to be higher than the experimental values in soft clay<br />

indicating the effect of soil <strong>com</strong>pressibility on the static penetration resistance.<br />

The ratios of predicted to measured penetration resistance are presented in<br />

table 2. It may be seen that these ratios are always greater than unity for<br />

Meyerhof's theory. Comparison with the theoretical values of Durgunoglu and Mitchell<br />

gives ratios which are close to unity for a dense deposit, indicating the validity of<br />

this method for general shear conditions. However, for low densities, these ratios<br />

are larger than one, indicating the significant influence of soil <strong>com</strong>pressibility on<br />

penetration resistance. The use of bearing capacity factors formulated for general<br />

shear conditions will cause overestimation of the penetration resistance of<br />

<strong>com</strong>pressible deposits.<br />

721


In <strong>com</strong>pressible soils, the shesr surface is restricted to s smaller zone around<br />

the penetrometer tip as suggested by Vesic (1963) for punching or local shear<br />

failures.<br />

CONCLUSIONS<br />

Among the various theoretical bearing capacity solutions available for the anal­<br />

ysis of the penetrometer problem, only those of Meyerhof (1961), Nowatzik and<br />

Karafiath (1978) and Durgunoglu and Mitchell (1975) take into account the penetrometer<br />

roughness, base apex angle and relative depth all of which are factors influencing the<br />

penetration resistance.<br />

As was seen in this analysis, Meyerhof' s method overestimates the penetration<br />

resistance in the deep foundation zone and is conservative in the shallow foundation<br />

region, while the method of Nawatzki and Karafiath is always conservative. The agree­<br />

ment between measured and predicted values using the theory of Durgunoglu and Mitchell<br />

for cohesionless and cohesion soils was reasonably good. This method can be used for<br />

predicting the static penetration resistance of relatively im<strong>com</strong>pressible soils.<br />

ACKNOWLEDGEMENTS<br />

The authors wish to thank Professor C.D. diCenzo, Dean of Engineering, for his<br />

constant encouragement and support of the research project. Thanks are also due to<br />

Dr. G.R. Peters, Associate Dean and Group Leader of Ocean Engineering for his con­<br />

structive <strong>com</strong>ments at various stages of the project. The assistance of Professor W.G.<br />

Smith in the development of the hardware is acknowledged with gratitude. Funding for<br />

the project is provided by a grant A-3710 from the Natural Sciences and Engineering<br />

Research Council Canada.<br />

NOMENCLATURE<br />

B: Width of foundation; diameter of penetrometer<br />

c: Soil cohesion<br />

D: Depth of foundation; depth of penetration<br />

I: Density Index (Relative density of sand)<br />

qf: Unit penetration resistance kPa<br />

w: Water content (moistrue content) %<br />

723


a: Semi Apex angle of cone<br />

Y dry : Dry density of soil<br />

6: Soil/foundation or soi1/pentrometer friction<br />

+: Angle of soil shear resistance<br />

REFERENCES<br />

Abde1-Gawad, S.M. (1979) "Static Penetration Resistance of Soils", M.Eng. Thesis,<br />

Memorial University of Newfoundland, St. John's, Nf1d., 225 p.<br />

Biarez, J., Burel, M. and Wack, B. (1961), "Contribution a l'etude de 1a force<br />

portante des foundations", <strong>Proceedings</strong> of the 5th International Conference on Soil<br />

Mechanics and Foundation Engineering, Paris, pp. 603-609.<br />

Chari, T.R., Smith, W.G. and Zielinski, A. (1978), "Use of Free Fall Penetrometer in<br />

Sea Floor Engineering", Conference Record, Ocean 78, IEEE-MTS Conference, pp. 686-<br />

691.<br />

Chari, T.R., Muthukrishnaiah, K. and Zielinski, A. (1979), "Performance Evaluation of<br />

a Free Fall Penetrometer", First Canadian Conference on Marine Geotechnical<br />

Engineering, Calgary, Alberta, April 1979, pp. 203-210.<br />

Chaudhuri, S.N. (1979), "Free Fall Impact Penetration Tests on Soils", M.Eng. Thesis,<br />

Memorial University of Newfoundland, St. John's, Nf1d., 134 p.<br />

De Beer, E.E. (1945), "Etudes des Fonditions sur Pilotis et des Fonditions Directes",<br />

Anna1es Des Travaux Publics de Belgique 46, pp. 1-78.<br />

De Ruiter, J. (1975), "The Use of In-Situ Testing for North Sea Soil Studies",<br />

Preprints, Offshore Europe 75, Aberdeen, pp. 219.1-219.10.<br />

Durgunog1u, H.T. and Mitchell, J.K. (1973), "Static Penetration Resistance of Soils",<br />

Research Report Prepared for NASA Headquarters, Washington, D.C., April 1973,<br />

University of California, Berkeley.<br />

Durgunog1u, H.T. and Mitchell, J.K. (1975), "Static Penetration Resistance of SOils,<br />

I-Analysis, II-Evaluation", <strong>Proceedings</strong> of the Conference on In-Situ Measurement of<br />

Soil Properties, ASCE, Vol. I, pp. 172-189.<br />

ESOPT, (1974), European Symposium on Penetration Testing, Stockholm, June 1974, Vol.<br />

I, State-Of-The-Art Report.<br />

Ferguson, G.H., McClelland, B. and Bell, W.D. (1977), "Seafloor Cone Penetrometer for<br />

the Deep Penetration Measurements of Ocean Sediment Strength", The 9th Offshore<br />

Technology Conference, OTC 2787, Vol. I, pp. 471-478.<br />

Hu, G.C. (1965), "Bearing Capacity of Foundation with Overburden Shear", Sols- SOils,<br />

Vol. I, No. 13, June 1965, pp. 11-18.<br />

Meyerhof, G.G. (1951), "The Ultimate Bearing Capacity of Foundations", Geotechnique,<br />

Vol. 2, pp. 301-322.<br />

724


Meyerhof, G.G. (1961), "The Ultimate Bearing Capacity of Wedge-Shaped Foundations",<br />

<strong>Proceedings</strong>, 5th International Conference on Soil Mechanics and Foundation<br />

Engineering, Vol. 2, pp. 105-109.<br />

Nowatzki, E.A. and Karafiath, L.L. (1972), "The Effect of Cone Angle on Penetration<br />

Resistance", Highway Research Board, No. 405, pp. 51-59.<br />

Nowatzki, E.A. and Karafiath, L.L. (1978), "Soil Mechanics for Offroad Vehicle<br />

ngineering", Series on Rock and Soil Mechanics, Vol. 2 (1974/77), No.5, Trans. Tech.<br />

Publications.<br />

Sanglerat, G. (1972), "The Penetrometer and Soil Exploration", Developments in<br />

Geotechnical Engineering, Elsevier Publishing Company.<br />

Terzaghi, K. (1943), "Theoretical Soil Mechanics", (Ninth Printing), John Wiley & Sons<br />

Inc., New York.<br />

Vesic, A.S. (1963), "Bearing Capacity of Deep Foundations in Sand", Highway Research<br />

Board Record, No. 39, pp. 112-135.<br />

Zuidberg, H.M. (1975), "The Sea Calf, a Submersible Cone Penetrometer Rig", Marine<br />

Geotechnology, Vol. 1, No.1, pp. 15-32.<br />

725


Hamdy Youssef<br />

Roger Kuhlemeyer<br />

ABSTRACl'<br />

DYNAMIC AND STATIC CREEP TESTING<br />

OF ICE AND FROZEN SOILS<br />

Genie Civil. Ecole Polytechnique de Montreal<br />

The University of Montreal. Quebec<br />

Department of Civil Engineering<br />

The University of Calgary. Alberta<br />

Canada<br />

Canada<br />

The behavior of frozen ground under dynamic loading has received very little<br />

attention to date. so that design information related to machine foundations in<br />

cold regions still very scarce. The objective of this paper is to demonstrate<br />

the importance of predicting the creep response of frozen ground (ice and frozen<br />

soils) to dynamic loadings; to present the state of knowledge in this area and<br />

to suggest a testing technique for investigation of the effect of the torsional<br />

vibration on the static creep rate of ice and frozen soils. as well as determi­<br />

ning their dynamic properties. this testing technique has been developed by the<br />

authors at the University of Calgary. Alberta. Canada.<br />

INTRODUCTION<br />

Frozen grounds can be defined as that half space in which stresses and strains<br />

arising due to the influence of an external load. are not constant; but vary with<br />

time. giving rise to a relaxation of stresses and creep (an increase in the strain<br />

with the passage of time). The main cause of the rheological processes in frozen<br />

soils is particulary due to their internal bonds. in which ice. which is an ideal­<br />

ly flowing solid plays a major role. (Tsytovich 1975).<br />

According to Ladanyi (1974). a basic foundation design philosophy in permafrost<br />

is based on predicting the delayed and long-term capacity of the foundation and its<br />

time dependent settlement due to <strong>com</strong>bined creep and consolidation. Predicting of<br />

time-dependent bearing capacity of frozen soils supporting static loading founda­<br />

tions is relatively easy. provided the creep and creep failure properties of the<br />

soil are known. On the other hand as the prediction of the delayed settlement is<br />

concerned. the frozen soil is considered as a quasi-single phase medium with ma-<br />

726


thematically well defined creep properties, neglecting the fact that one portion of<br />

the observed creep is actuallY due to vOlumetric strain (Ladanyi 1974, 1981).<br />

In projects involving construction of machine foundations in cold regions, the<br />

sub-and-super structures will be constructed first and then the machines (e.g. ge­<br />

nerators and <strong>com</strong>pressors) will be situated in place. These <strong>com</strong>ponents will deter­<br />

mine the magnitude of the static load which the frozen ground has to support over<br />

a long term, usuallY the life time of the structure.<br />

PREDICTION OF CREEP DEFORMATION DUE TO STATIC LOAD (STATIC CREEP)<br />

Under this static load, instantaneously, elastic deformation (i.e. strain) will<br />

occur, and with time continuous deformation (creep) will progress; firstlY at a de­<br />

creasing rate (stage I - Fig. 1) then followed by a steady state creep (stage II -<br />

Fig. 1). The long-term strength is represented by the end of this stage (point B -<br />

Fig. 1), the load has to be designed so that creep failure (accelerating creep -<br />

stage III) will not occur.<br />

Ladanyi (1972) developed a unified engineering theory of time, temperature and<br />

normal pressure dependent deformation and strength of frozen soils, which enabled<br />

the creep and creep failure information to be expressed in a relativelY simple ana­<br />

lYtical form, using a minimum number of experimental parameters. This theory has<br />

been highly credited and widelY adopted in recent years by researchers and practi­<br />

ce engineers for prediction the static creep of frozen ground. More recently, Li<br />

and Andersland (1980) investigated experimently the effect of cyclic loadings on<br />

the creep parameters in Ladanyi's theory and they found that the cyclic load acce­<br />

lerate the static creep rate, this effects appear to be included in the creep para­<br />

meter n and the proof stress 0c'<br />

Ladanyi (1972) assumed that for long-term deformations, the frozen soil behavior<br />

is generallY dominated by a secondary creep, and the creep curve could be correctlY<br />

approximated by a straight line having the intercept E(i) at time t - 0, as shown in<br />

Fig. 1. Therefore, for a constant stress and temperature the total strain E can be<br />

predicted as:<br />

where:<br />

E = Young's modulus, ° = uniaxial normal stress, Ek = arbitrary small strain, Ok =<br />

proof stress for E k , Ec = arbitrary creep rate, 0c = proof stress for Ec' and k,n =<br />

creep parameters.<br />

727


3. Response Prediction<br />

The analysis of the response of frozen ground to dynamic loadings requires: (1)<br />

determining appropriate material properties (which are available) and (2) selecting<br />

a suitable analytical techniques to predict ground response to machine foundation<br />

(Vinson 1978). As previously mentioned these analytical techniques do not exist as<br />

yet, and the designer has to use the dynamic soil properties of the frozen ground as<br />

an input to suitable analytical techniques developed for unfrozen soils with great<br />

careto insure that the assumptions on which the techniques are based are not violated.<br />

Due to the fast development of Northern Canada and Alaska; the joint U.S.-Canadian<br />

<strong>com</strong>mittee in the Northern Civil Engineering Research workshop (Carlson et al. 1978)<br />

stated that there is a need for research in: (1) basic studies on the interaction<br />

between high-frequency dynamic foundation loads and frozen ground, and (2) reduction<br />

in shear and adfreeze strength and possible acceleration of creep caused by vibratory<br />

loadings.<br />

The above discussion aimed to place emphasis on the necessity for analytical tech­<br />

niques which should be developed specifically for predicting the response of frozen<br />

ground to dynamic loadings and especially for the dynamic creep. These analytical<br />

techniques should be based on and verified by experimental and field data, which are<br />

lacking. Therefore, the first step to acheive this goal is the development of expe­<br />

rimental apparatus to provide this needed information. The ideal apparatus should be<br />

capable of permitting investigation the influence of the following parameters: (1)<br />

temperature, (2) strain or deformation with time, (3) axial and confining stresses,<br />

(4) frequency and amplitude for longitudinal and torsional vibration, (5) different<br />

material type and <strong>com</strong>position, and (6) dynamic properties during different stages of<br />

the test. The authors have developed an experimental apparatus at the University of<br />

Calgary, Alberta, Canada to investigate most of the above mentioned parameters, a<br />

breif description of Calgary apparatus is given below.<br />

4. Calgary Apparatus<br />

In connection with the design of the proposed Mackenzie Valley pipe line, this ap­<br />

paratus has been developed to study in the laboratory the effect of vibration on the<br />

static creep rate of ice and frozen soils as well as their dynamic properties.<br />

The new (1976) Drnevich free-free torsional resonant column apparatus has been<br />

used as a forced vibration device. This device has been modified at the University<br />

of Calgary to permit applying vertical loads without influencing the free torsional<br />

vibration characteristics of the sample apparatus system. The testing procedures<br />

which are to be followed and which the apparatus is designed for is according to the<br />

line of thought as described before and expressed in Fig. 1. Two vertical identical<br />

730


Fig. 2. CALGARY APPARATUS<br />

SUGGESTED TESTING TECHNIQUE FOR EXPERIMENTAL<br />

INVESTIGATION OF STATIC AND DYNAMIC CREEP, AS WELL AS<br />

THE DYNAMIC PROPERTIES OF ICE AND FROZEN SOILS .<br />

cylindrical samples of ice or frozen soils are subjected to the same constant tempe­<br />

rature and vertical static loading conditions. After time tsd (Fig. 1) the two spe­<br />

cimens will strain to Esd. At this time a steady- state sinusoidal torsional vibra­<br />

tion is superimposed to one specimen (in the left triaxial cell in Fig. 2) while the<br />

other specimen is subjected to static loading only. The torsional vibration is ap­<br />

plied to the bottom end of the dynamically tested specimen while the top end is free<br />

(to r otate) except for a light , relatively rigid cap ; the input motion and output<br />

sample r esonance are both observed and measured with piezoelectric t r ansducers atta­<br />

ched to the cap and base plates of the specimen. The axial deformati ons of the two<br />

specimens are recorded with time , which allow obtaining the static and dynamic creep<br />

curves as shown in Fig . 1 ; for di ffer ent types of ice and frozen soil s . The dynamic<br />

properties can be measured and evaluated during different stages of the test from<br />

the resonant column theory (Drnevich 1976a) and by using Drnevich (1976b , 1978) reso-<br />

731


ACKNOWLEDGEMENTS<br />

The authors wish to acknowledge the work contribution by Mr. R. French, of Geo­<br />

Phsi-Co. Ltd - Calgary, Alberta, during the initial development of Calgary Appara­<br />

tus. Financial support from the University of Calgary and the National Research<br />

Council of Canada is gratefully acknowledged. The authors are grateful to Prof.<br />

B. Ladanyi of Ecole Polytechnique de Montreal for the help provided during prepara­<br />

tion of this paper.<br />

REFERENCES<br />

(1) ANDERSLAND, 0., SAYLES, F. and LADANYI, B. (1978), "Mechanical Properties of<br />

Frozen Ground", Chp. 5 in "Geotechnical Engineering :for Cold Regions", Mc<br />

Graw-Hill Book Co., N.Y., U.S.A.<br />

(2) BARKAN, D. (1962), "Dynamics of Bases and Foundations", McGraw-Hill Co., N.Y.<br />

(3) CARLSON, R. and MORGENSTERN, N. (1978), "Northern Civil Engineering Research<br />

Workshop Report", Editor, The Univ. of Alberta, Edmonton, March 20, 1918.<br />

(4) DRNEVICH, V.P. (1976a), "Free-Free Resonant Column Apparatus Operating Manual",<br />

Soil Dyn. Inst. Inc., Lexington, Kentucky, U.S.A.<br />

(5) DRNEVICH, V.P. (1916b), "A user's Manual for the FORTRAN Computer Program En­<br />

titled RESCOL4", Dept. of civil Engg., Univ. of Kentucky, U.S.A.<br />

(6) DRNEVICH, V.P. (1978), "Resonant Column Test", Report No. S-78-6, Geotech. Lab.<br />

U.S. Army WES, Vicksburg, Miss., U.S.A.<br />

(7) IVANOV, P. L. (1980), "Vibrocreep and Loose soil strength under cyclic loading<br />

Action", Int. Symp. on Soils Under Cyclic and Trans. Ldg., Swansea, U.K.<br />

(8) LADANYI, B. (1972), "An Engineering Theory of Creep of Frozen SoilS", Canadian<br />

Geot. J., 9, 63.<br />

(9) LADANYI, B. (1914), "Bearing Capacity o:f Frozen Soils", Proc. 21th Canadian<br />

Geot. Conf., Edmonton, Canada.<br />

(10) LADANYI, B. (1981), "Mechanical Behavior of Frozen Soils", Proc. Int. Symp. on<br />

the Mech. Behav. of Structured Media, Carleton Univ., Ottawa, Canada.<br />

(11) LI, J. and ANDERSLAND, o. (1980), "Creep Behavior of Frozen Sands Under Cyclic<br />

Loading Conditions", Proc. 2nd Int. Symp. on Ground Freezing, Trondheim,<br />

Norway.<br />

(12) PANDE, G. and SHARMA, K. (1980), "A Micro-Structural Model for Soils Under<br />

Cyclic Loading", Proc. Int. Symp. on Soils Under Cyclic and Tran. Ldg.,<br />

Swansea, England.<br />

(13) SAYLES, F. (1974), "Triaxial Constant Strain Rate Tests and Triaxial Creep<br />

Tests on Frozen ottawa Sand", U.S. Army CRREL, Hanover, N.H., U.S.A.<br />

(14) SHETTY, D., MURA, and MESH II (1915), "Analysis of Creep Deformation Under Cyclic<br />

loading Conditions", Material Science and Engg., 20-261-266, Elsevier Seq.,<br />

Netherlands.<br />

733


(15) STEVENS, H. (1975), "The Response of Frozen Soils to Vibratory Loads", U.S.<br />

Army CRREL, Tech. Rep. 265, Hanover, N.H., U.S.A.<br />

(16) TRIMBLE, J.R. (1977), "A Comparison of the Creep Deformation on Naturally<br />

Frozen Soils Under Static and Repeated Loadings", M.Sc. Thesis, Queen's<br />

Uni v., Canada.<br />

(17) TSYTOVICH, N. (1975), "The Mechanics of Frozen Ground", McGraw-Hill Co., N.Y.<br />

(18) TURCOTT-RIDS (1980), "Resonant Column Testing of Frozen Soils", M.Sc. Thesis,<br />

McGill Univ., Montreal, Canada.<br />

(19) VINSON, T. (1978), "Response of Frozen Ground to Dynamic Loading", Chap. 8,<br />

in "Geotechnical Engg. for Cold Regions", McGraw-Hill Co., N.Y., U.S.A.<br />

(20) YOUSSEF, Hamdy (1979), "Development of a Testing Apparatus for Static and<br />

734<br />

Dynamic Creep Testing of Ice and Frozen Soils", M.Sc. Thesis, The Univ.<br />

of Calgary, Alberta, Canada.


William M. Sackinger<br />

Associate Professor<br />

Abstract<br />

A Review of Technology for<br />

Alaskan Offshore Petroleum Recovery<br />

Geophysical Institute<br />

University of Alaska<br />

Fairbanks, Alaska 99701<br />

USA<br />

A <strong>com</strong>prehensive summary is presented of all of the environmental<br />

hazards which relate to the design and deployment of petroleum produc­<br />

tion systems for the Alaskan Beaufort, Chukchi, and Bering Seas.<br />

Environmental factors which control the design choices are identified for<br />

each area, and the most recently available values for the magnitude of<br />

these hazards are discussed. Both structure designs and icebreaking<br />

vessel operations are shown to be dependent upon these selected para­<br />

meters. Design example calculations are shown to be dependent upon<br />

these selected parameters. Design example calculations are made for<br />

structures in the Bering, Chukchi, and Beaufort offshore lease areas.<br />

Finally, research which would lead to optimization of such structures<br />

is suggested.<br />

Introduction<br />

The offshore regions of Alaska, most of which are subject to<br />

moving sea ice for at least part of the year, have be<strong>com</strong>e prominent as<br />

major unexplored areas from which future domestic petroleum may be<br />

produced. In Figure 1, the geologic basins of the Bering, Chukchi,<br />

and Beaufort Seas are broadly illustrated. In the Beaufort Sea, many<br />

areas are within reasonable distance of the trans-Alaskan pipeline's<br />

northern terminal; other more remote regions in the Chukchi and<br />

Bering Seas do not have ready access to an existing transportation link,<br />

735


736<br />

implying that marine transportation (including pipelines and/or<br />

tanker loading facilities) will be required if oil is discovered in<br />

these ice-infested seas.<br />

Production structures, oil gathering and storage facilities,<br />

tanker loading terminals, ice-breaking tankers, and ice-breaking<br />

support vessels will be required. Designs will often be domina­<br />

ted by ice considerations, but in some regions (such as the<br />

southern Bering Sea) other environmental hazards such as waves,<br />

earthquakes, and seafloor sediment conditions must also be taken<br />

into account, and in fact may dictate the final designs. In this<br />

paper, the important environmental hazards for each area are iden­<br />

tified and quantified, to the extent permitted by the present<br />

in<strong>com</strong>plete state of knowledge of these remote regions, and some<br />

sample calculations of ice forces on structures are made for several<br />

ice-dominated areas. In assembling this information, reference is<br />

made to the results of many researchers, and much of their work<br />

has appeared in previous POAC Conferences. Finally, suggestions<br />

are made for future research which would make offshore produc­<br />

tion from these areas of severe ice problems technologically prac­<br />

tical.<br />

Environmental Hazards<br />

In the course of structure design, the meteorological vari­<br />

ables of wind, temperature and pressure are, of themselves, of<br />

little direct consequence, but they are responsible for waves,<br />

storm tides, ocean currents, structure ice accretion, and sea ice<br />

movement, all of which must be included in structure design calcu­<br />

lations. In Figure 2, the inter-relationships of these n secondary"<br />

environmental forces are shown. Bathymetry is an important factor<br />

in assessing the feasibility of several types of offshore structures<br />

for Alaskan waters. In Table I, the bathymetry range for the<br />

several basins is given. For shallow water, up to 20 meters in<br />

depth, the artificial island has proven to be useful in the Canadian<br />

and Alaskan Beaufort Sea. The caisson-retained artificial fill island<br />

concept might be extended to greater depths, such as 30-40 meters,<br />

but the cost of such an approach would have to be weighed against


other options such as conical gravity structures. For deeper water<br />

in the Chukchi Sea (40 meters) these latter appear promising, and<br />

for the 100 to 150 m depths of the Bering Sea, steel or concrete<br />

slender structures are being considered.<br />

Wave-loading and current-loading calculations have be<strong>com</strong>e rather<br />

refined and well-understood for offshore platforms in non-ice-covered<br />

areas, and will not be repeated here. The impact of an isolated ice<br />

floe driven by high waves against the leg of an offshore platform has<br />

not yet been carefully analyzed, however.<br />

A widely-quoted source of information for maximum wave heights<br />

in the Alaskan offshore areas is the Climatic Atlas of the Outer Contin­<br />

ental Shelf Waters and Coastal Regions of Alaska [II. For the several<br />

regions of the Bering Sea, Figure 3 shows the wave heights which are<br />

predicted, based on calculations using the approach of Thom [2,3 I.<br />

Such calculations, however, do not include the possibility that wind<br />

fetch and also wave height are reduced by the possible presence of an<br />

ice cover. Furthermore, shallow-water effects (which would be expec­<br />

ted in Norton Sound [20 meters nominal depth]) and non-linear wave<br />

effects are not included. Extreme waves of 40 meters for a 100-year<br />

return period are difficult to imagine in the Bering Sea, but clearly<br />

the establishment of a realistic extreme wave height is of importance<br />

for the Navarin, St. George, Zemchug, Bristol and St. Matthew Basin<br />

regions, which are frequently exposed directly to weather systems<br />

from the southwest.<br />

Ice accretion on offshore structure and service vessels can also<br />

be a serious hazard, greatly increasing their weight and elevating the<br />

center of gravity. Very large loads can accumulate; for example,<br />

drilling operations in Lower Cook Inlet were interrupted when a semi­<br />

submersible rig accumulated an estimated 450 metric tons of vertical<br />

load due to ice during winter operations. Freezing spray <strong>com</strong>monly<br />

occurs [1] when the air temperature is between -2°C and -18°C, the<br />

sea surface temperature is below +5 0 C, and the wind speed is greater<br />

than about 11 m/sec. Using weather data from St. Paul [II, in the<br />

Pribilov Islands adjacent to the St. George Basin of the Bering Sea,<br />

one finds for example that in January the air temperature is less than<br />

737


738<br />

-8°C and the average wind speed is greater than 11 meters/sec. for 9%<br />

of the time. The sea surface temperature in that region is less than<br />

+loC for 10% of the time. For the <strong>com</strong>bination of these conditions, spray<br />

icing will accumulate at rates from 3.5 cm to over 14 cm thickness in 24<br />

hours [1). Although exploratory drilling in summer in that area would<br />

not encounter this hazard, winter drilling or year-round production<br />

must take superstructure icing into account. Designs could allow for<br />

this additional gravity and wind loading, or, alternatively, special<br />

surface coatings with low adfreeze bonding (such as PTFE) could be<br />

used (4). Waste heat and special y designed enclosures rna y also be<br />

utilized. An additional type of ice accumulation can be expected from<br />

floating frazil ice particles (5) which adhere to the legs of a structure<br />

near the waterline. The adhesion rate is a function of the turbulence<br />

around the structure legs, and in protected Arctic harbors, has been<br />

related to the tidal cycle. Further research is needed to determined the<br />

degree of frazil ice adhesion which would be encountered under open<br />

ocean conditions in the marginal ice zone.<br />

There are, broadly speaking, four categories of moving ice hazards,<br />

in order of increasing severity: the undeformed sheet ice; annual ice<br />

ridges; multi-year ridges; and ice islands. Each of these ice forms<br />

produces lateral and vertical loads on structures, which are determined<br />

by the ice thickness, the ice velocity, and the mechanical properties<br />

of the ice. The details of the several modes of ice/structure interac­<br />

tions will be mentioned in a later section of this paper.<br />

A convenient geographical division may be made at the Bering<br />

Strait, on the basis of ice severity. In the Bering Sea south of the<br />

Bering Strait, multi-year ice is rarely encountered, although for some<br />

winter weather conditions the sea ice (presumably including both annual<br />

and multi-year broken ice floes) does flow southward from the Chukchi<br />

Sea through the Strait (6). Additional field observations have been<br />

planned to assess this condition. In the Chukchi and Beaufort Seas<br />

north of the Bering Strait, multi-year ice is <strong>com</strong>monly encountered, but<br />

ice islands are found infrequently.<br />

Regardless of the choice of type of offshore structure, problems are<br />

associated with its installation. It is far easier to install most structures


742<br />

A <strong>com</strong>parison of this result for a multi-year ridge with that of<br />

annual ice in the Bering Sea shows that the lateral force is nearly 8 times<br />

greater in the Chukchi and Beaufort Sea pack ice region than in the most<br />

hazardous part of the Bering Sea. The <strong>com</strong>parison with sea ice in Cook<br />

Cook Inlet can also be made; assuming an ice thickness of 1.3 meters in<br />

Cook Inlet, and crushing failure, lateral forces of about 22,460 kN are<br />

predicted by using Korzhavin's formula and the same assumptions for<br />

other parameters as in the Bering Sea. The ratio of horizontal forces<br />

in the Chukchi Sea to those in Cook Inlet are, in this simplified example,<br />

61/1. Obviously, structures for the Chukchi Sea must be much more<br />

robust and have a very wide foundation to counteract the overturning<br />

moment. For discussion of these details, the reader is referred to the<br />

thesis by Karp [15 I.<br />

One final observation must be made. A search for a structure<br />

coating with low friction coefficient would appear to have a potentially<br />

dramatic benefit, reducing the calculated horizontal force to perhaps as<br />

low as 913 ,600 kN in the Chukchi Sea, or only 40 times greater than<br />

the Cook Inlet situation. Changes to lower cone angles would also be<br />

of some benefit. Further research on low-friction coatings for structures<br />

subject to moving ice loads is strongly re<strong>com</strong>mended.<br />

One could criticize the assumptions made in Karp's analysis, pointing<br />

out (as he did) (15) that multi-year keel depths of 52 meters probably<br />

exist, since scouring of the seafloor by ice has been observed to this<br />

depth. Moreover, the flexural strength of multi-year ice is known to be<br />

widely variable. Finally, ice islands have not yet been considered.<br />

These and related supporting research studies remain for the future [17 I.<br />

Geological Hazards<br />

A <strong>com</strong>prehensive book by G. D. Sharma [181 has recently been<br />

published which provides much useful information on sediment size<br />

distribution in the Alaskan offshore areas. For the St. George, Norton,<br />

and Beaufort Basins, research has revealed few seafloor problems that<br />

would be considered severe for offshore foundation design [171, with<br />

the notable exception of the very fine gas-charged sediments in Norton<br />

Sound, as described by Thor and Nelson [191. Formation and growth


of gas craters under gravity structures could occur there, if no provi­<br />

sion were made for normal gas evolution to the surface. Geological<br />

reconnaissance and geotechnical properties studies remain to be in the<br />

other areas.<br />

Earthquakes have been measured for many decades, and most earth­<br />

quakes occur along the Aleutian arc. Activity in St. George Basin and<br />

the Seward Peninsula seems to be limited to events of less than magnitude<br />

7.0 Richter. The transfer of energy from the epicenter to a particular<br />

site, and the seafloor accelerations at the site, remain to be determined,<br />

however. Seed (20) has studied liquefaction of marine soils due to earth""<br />

quakes and has concluded that for most cases an acceleration of the<br />

order of 0.16 g, and a magnitude 7.5 Richter, is required for soil lique­<br />

faction. This result has not been verified for gas-charged marine sedi­<br />

ments, however. Although foundation design and earthquake excitation<br />

of modes of vibration of the structure should be considered, these prob­<br />

lems are not peculiar to the Arctic, nor does the Arctic present any<br />

obvious problems, with the possible exceptions of the gas-charged sedi­<br />

ments of Norton Basin, and the presence of ice adjacent to vibrating<br />

structures during earthquakes.<br />

The offshore permafrost and in-situ gas hydrates of the Beaufort<br />

and Chukchi Seas will call for special drilling techniques and cementing<br />

procedures, which are, however, beyond the scope of this paper (17).<br />

Research Frontiers in the Alaskan Offshore Areas<br />

On 30 June - 2 July 1980, this writer convened a small research<br />

planning workshop, on behalf of the U.S. Department of Energy, at<br />

Sandia National Laboratories, Albuquerque, New Mexico. The report<br />

of the workshop (21) identified 60 engineering problem areas for further<br />

research; these are listed in the Appendix. The workshop was preceded<br />

by a more <strong>com</strong>prehensive review report (17), covering some of the<br />

material discussed in this paper, and also identifying research needs.<br />

Conclusions<br />

1. Although lateral forces caused by wave loading are likely to<br />

dominate over ice forces in the southern Bering Sea, an accur­<br />

ate assessment of extreme wave heights remains to be <strong>com</strong>pleted<br />

for the Bering Sea.<br />

743


744<br />

2. Spray ice buildup on structures, ships, and aircraft will occur for<br />

an appreciable percentage of the time during the winter months in<br />

the Bering Sea. This must be prevented or ac<strong>com</strong>modated with<br />

special designs.<br />

3. Rafted annual ice floes up to 10 m thick in the Bering Sea could<br />

produce lateral forces some eight times greater than those in Cook<br />

Inlet. The probability of encounter between these floes and a<br />

production structure is quite low but remains to be established.<br />

4. Multi-year ice ridges breaking against a conical structure in flexure<br />

represent a typical, severe ice loading condition in the pack ice of<br />

the Chukchi and Beaufort Seas. A sample calculation gives lateral<br />

forces ranging from 40 to 61 times larger than those due to ice<br />

crushing in Cook Inlet, depending upon the friction between ice<br />

and structure.<br />

5. Many technology research directions remain to be addressed in order<br />

to design and emplace safe and economical offshore production struc­<br />

tures in the pack ice of the Beaufort, Chukchi and Bering Seas.<br />

Acknowledgements<br />

Much of this material has been reviewed with the support of the<br />

U.S. Department of Energy, Division of Oil, Gas and Shale Technology.<br />

This study was supported by the Bureau of Land Management through<br />

interagency agreement with the National Oceanic and Atmospheric Adminis­<br />

tration, under which a multi-year program responding to needs of petro­<br />

leum development of the Alaskan Continental Shelf is managed by the<br />

Outer Continental Shelf Environmental Assessment Program Office. The<br />

review of this manuscript by G. Weller was most helpful.


746<br />

13. Schwarz, J., and W. F. Weeks (1977). "Engineering properties<br />

of sea ice", Journal of Glaciology, 19, pp. 499 531.<br />

14. Wang, Y. S. "Sea ice properties", Technical Seminar on Alaskan<br />

Beaufort Sea Gravel Island Design, Exxon USA, Anchorage,<br />

October IS, 1979.<br />

15. Karp, L. B. "Concept Development of a Concrete Structure<br />

Founded in the Ice-Stressed Chukchi Sea: A Case of Ice/Structure<br />

Interaction in an Offshore Arctic Region", D. Eng. thesis,<br />

University of California, Berkeley, August 1980.<br />

16. Frederking, R. "Dynamic ice forces on an inclined structure", in<br />

Physics and Mechanics of Ice, P. Tryde (ed.), Springer-Verlag,<br />

Berlin, 1980, pp. 104-116.<br />

17. Sackinger, W. M. "A review of technolo gy for arctic offshore oil<br />

and gas recovery", U. S. Department of Energy, Division of<br />

Fossil Fuel Extraction, Contract No. DE-A COI-80ET14317,<br />

Vol. I, June 6, 1980, 97pp.<br />

18. Sharma, G. D. The Alaskan Shelf, Springer-Verlag, New York,<br />

1979.<br />

19. Thor, D. R. and I-l. Nelson. "A summary of interacting, surficial<br />

geologic processes and potential geologic hazards in the Norton<br />

Basin, northern Bering Sea", Proc. of the Offshore Technology<br />

Conference, Houston, TX (1979), pp. 377-385, (OTC 3400).<br />

20. Seed, H. B. "Soil liquefaction and cyclic mobility evaluation for<br />

level ground during earthquakes", J. Geotechnical Eng. Div.,<br />

Proc. ASCE, Vol. 205, No. GT2, February 1979, pp. 201 255.<br />

21. Sackinger, W. M. "Report of the workshop on arctic oil and gas<br />

recovery", U.S. Department of Energy, Office of Oil, Final<br />

Report for Contract No. DE-ACOI-80ET14317, September 1980,<br />

38pp.


Appendix I<br />

Re<strong>com</strong>mended research topics in support of Arctic offshore petroleum<br />

recovery: The results of a workshop at Sandia National Laboratories,<br />

Albuquerque, New Mexico, 30 June - 2 July 1980.<br />

1. Verify hindcast models for Arctic areas that include the influence<br />

of ice cover on wave and surge generation, and shallow water<br />

effects.<br />

2. Investigate methods to date ice gouges of the seafloor.<br />

3. Investigate mechanics of ice gouging as a function of sediment<br />

type.<br />

4. Conduct confined and unconfined tests to provide data on strength<br />

and stress/strain behavior of sea ice and freshwater ice as a<br />

function of environmental parameters and loading rates.<br />

5. Conduct in-situ beam tests to obtain flexural strength data.<br />

6. Conduct large scale and small scale tests on the same sea ice to<br />

investigate size effects.<br />

7. Investigate analytical procedures to predict size effects.<br />

8. Obtain data on dielectric and electromagnetic properties of sea ice.<br />

9. Obtain engineering data on frictional forces between ice and other<br />

materials.<br />

10. Obtain data on adhesion bond strength.<br />

11. Collect in-situ measurements of first-year and multi-year ridge<br />

geometry and properties.<br />

12. Measure ice drift velocity.<br />

13. Develop and verify regional ice dynamics models.<br />

14. Conduct systematic stereo photography, laser overflights, and<br />

sonar measurement programs to obtain ridge profile distributions.<br />

15. Verify use of passive microwave imagery to identify multi-year ice.<br />

16. Process available satellite imagery to provide data on multi-year<br />

ice movement and floe size distribution.<br />

17. Develop physical model of pack ice edge movement.<br />

747


748<br />

18. Compile pack ice edge data to verify the ice edge movement model.<br />

19. Obtain field measurements of ice state, e.g., crystal structure,<br />

<strong>com</strong>position, temperature, and snow cover.<br />

20. Utilize ships of opportunity and specific icebreaker voyages to<br />

collect data on ice conditions.<br />

21. Conduct oceanographic/meteorological measurement program.<br />

22. Conduct systematic side-scan sonar and bathymetric surveys in<br />

the Beaufort Sea to water depths of 80 m and in the Chukchi Sea.<br />

23. Sponsor a logistics base at the Naval Arctic Research Laboratory,<br />

Barrow, and an equivalent operating base at Nome to conduct<br />

field studies in the Chukchi and Bering Seas.<br />

24. Improve understanding of failure modes of a variety of ice features<br />

against various types of structures.<br />

25. Investigate the influence of random flaws and nonsimultaneous<br />

structure contact on total ice forces.<br />

26. Develop a model of driving force and average ridge-building<br />

forces across a wider structure front.<br />

27. Describe local ice pressures as a function of ice mechanical<br />

properties, ice velocity, and contact geometry.<br />

28. Investigate the characteristics and effects of rubble fields around<br />

structures.<br />

29. Consider test structure verification of ice/structure interaction<br />

models.<br />

30. Conduct model tests and develop predictive theories for ice<br />

rideup on islands and structures.<br />

31. Develop improved techniques to predict wave runup and overtopping.<br />

32. Investigate non-linear surface effects for wave diffraction theory.<br />

33. Evaluate impact loads upon structures from storm-driven ice floes.<br />

34. Conduct studies to investigate ice actions on slope protection<br />

systems.<br />

35. Improve prediction of ice ridge penetration resistance for icebreakers.<br />

36. Develop models of ship/multiyear ice ridge breaching.


37. Develop model to predict ice pressure-induced delays in ship<br />

transit.<br />

38. Deveop a verified model for high energy penetration, resistance,<br />

and pressure on a ship's hull pursuant to collision with an ice<br />

feature.<br />

39. Accelerate development of models to predict ice forces on propellers<br />

and conduct full scale verification trials.<br />

40. Conduct full scale and model tests to improve propulsion efficiency<br />

in ice.<br />

41.<br />

42.<br />

43.<br />

44.<br />

45.<br />

46.<br />

47.<br />

48.<br />

49.<br />

50.<br />

51.<br />

52.<br />

Extend existing models of ship/terminal interactions in ice to more<br />

severe ice conditions of the Arctic offshore. Verify with model<br />

tests.<br />

Develop mathematical models and physical modeling techniques to<br />

improve ship maneuvering performance in ice.<br />

Develop models to predict variations in ship performance in level<br />

ice due to changes in ice strength, hull shape, and snow cover.<br />

The characterization of the seafloor should be <strong>com</strong>pleted in the<br />

regions of the Arctic offshore which will be subject to petroleum<br />

develop men t •<br />

The mechanical properties of sediments on the sea bed should be<br />

determined.<br />

Geomorphological processes acting upon the seafloor, such as ice<br />

scour, must be described more precisely •<br />

The ground motion offshore • which results from earthquakes should<br />

be measured.<br />

A model of thaw and subsidence for a warm subsea pipeline buried<br />

in offshore permafrost should be developed.<br />

Frost heave in marine soils and gravels containing seawater<br />

deserves further investigation.<br />

The procedures for predicting behavior of pilings and island<br />

foundations installed above subsea permafrost should be refined.<br />

Effects of artificial islands and causeways upon natural coastal<br />

processes should continue to be studied.<br />

The location and extent of ice-bonded subsea permafrost should<br />

be determined for the Arctic offshore basins which are expected<br />

to be developed.<br />

749


750<br />

53. A remote sensing technique for detection of gas hydrates, in-situ,<br />

should be perfected.<br />

54. The location of active subsurface faults should be determined.<br />

55. Both theoretical and remote-sensing techniques should be applied<br />

to ascertain the extent of thaw of gas hydrates during production,<br />

and the possible resulting subsidence.<br />

56. The technology of cementing casing in regions where gas hydrates<br />

have been penetrated should be perfected.<br />

57. Instrumentation for both laboratory and in-situ measurements of<br />

ice properties should be reviewed and perfected.<br />

58. Methods should be developed for rapid and easy assessment of<br />

ice ridge profiles and degree of consolidation.<br />

59. Diagnostic instrumentation should be developed to measure the<br />

interaction of ice with prototype or operational artificial islands,<br />

structures, and subsea pipelines.<br />

60. An advanced ice surveillance system to be used in an operational<br />

mode for ship routing in ice-infested waters should be designed.


FIG. 1. Selected Arctic Alaskan Offshore Basins.


IG. 2. Major Technical Factors Related to Arctic Offshore Oil and Gas Recovery<br />

LAND<br />

PIPELINES<br />

IN<br />

PERMAFROST


R. G. Sisodiya<br />

K. D. Vaudrey*<br />

Abstract<br />

BEAUFORT SEA FIRST-YEAR<br />

ICE FEA TURES SURVEY - 1979<br />

Gulf Research and<br />

Development Co.<br />

Vaudrey & Associates<br />

Houston, TX, USA<br />

Missouri City, TX, USA<br />

A joint industry study was conducted on first-year ice features in the Alaskan Beaufort<br />

Sea during March-April 1979 to determine ice feature geometry and internal characteristics as<br />

well as assess winter ice conditions in the lease sale area offshore Prudhoe Bay.<br />

Data from the field investigation consisted of sail and keel profiling of ridges by standard land<br />

surveying and sonar techniques, respectively, while internal <strong>com</strong>position was defined by ice<br />

augering.<br />

Over 300 miles of stereo aerial photography was flown both perpendicular and parallel to the<br />

coast outside of the barrier island chain •.<br />

I. Introduction<br />

Since first-year ice ridges and grounded features form each winter in the Alaskan Beaufort Sea,<br />

they may pose an operational problem for transportation and pipeline installations. In some<br />

instances, a consolidated first-year ridge moving against a structure may provide the design<br />

load.<br />

Several joint industry projects have utilized aerial photography and laser profilometry to<br />

determine first-year ridge populations; but data are limited on geometry statistics and<br />

consolidation properties. Only a few individual first-year ridges have been investigated (I), (2),<br />

(3), (4). Ridging frequency data have been determined for large offshore areas in the Alaskan<br />

Beaufort Sea (4), (5), (6), (7), but may not be applicable for the nearshore area of interest.<br />

*Formerly with Gulf Research and Development Co.<br />

755


Gulf Research and Development Company, along with seven participating <strong>com</strong>panies, performed<br />

a joint industry study to investigate first-year ice features in the Prudhoe Bay region of the<br />

Alaskan Beaufort Sea during late winter 1979. The objectives of this program were: I) to<br />

provide statistical information on ice features that may be correlated to predict ridge-structure<br />

interaction; 2) to determine ridge and rubble pile geometries to help establish pileup and<br />

loading design criteria; and 3) to determine internal ridge characteristics to provide insight on<br />

possible models of first-year ridge behavior.<br />

2. Field Investigation<br />

An on-the-ice survey of thirteen first-year ridges and two rubble piles was performed,<br />

determining sail profiles by standard surveying, keel profiles by sonar techniques (fathameter<br />

with 3 0 transducer), and internal <strong>com</strong>position by ice augering. Ice features studied are shown<br />

on the map in Figure I. Twenty-two <strong>com</strong>plete cross-sections were obtained with seven of the<br />

ridges having two or more profiles eoch.<br />

The average keel depth to sail height (K/S) ratio for 17 floating ridge sections was calculated as<br />

K/S = 5.5 ! 1.2. All ridge sections with sail heights greater than 15 feet were grounded. The<br />

average K/S ratio lies between the results of Kan (8), who found an average K/S = 7.6 for five<br />

ridges, and Kovacs (I), who calculated K/S = 4.9 for five profiles from three separate ridges.<br />

For floating ridges 113-12 the sail height is plotted versus keel depth in Figure 2, showing a well­<br />

correlated linear relationship given by a least squares straight line fit of the data points.<br />

Assuming that keel depths are distributed normally about the mean defined by the straight line,<br />

95% of the keel depth observations should lie within the confidence intervals shown in Figure 2<br />

as dashed lines.<br />

A method using the program data was established for determining a critical sail height (Scr) for<br />

a given water depth, representing a boundary between floating and grounded ridges. The<br />

critical sail height can be plotted as a function of water depth as shown in Figure 3 as a straight<br />

line least squares fit minimizing sail height error.<br />

At least two augerholes were drilled into the keel of each ridge section to determine internal<br />

<strong>com</strong>position, and in most cases, solid, <strong>com</strong>petent ice existed up to 6-8 feet below sea level,<br />

indicating that the keels were not very well-bonded below a single, late-winter ice sheet<br />

thickness. Where rafting occurs extensively, free water was generally absent at the ice sheet<br />

layer boundaries. Therefore, large refrozen ice floes approximately 18-24 feet thick may exist<br />

between and on the slopes of ridges. For example, Ridge 9 shown in Figure 4 had a rafted floe<br />

along its southern boundary greater than 22 feet thick, indicating at least four five-foot ice<br />

756


sheet layers rafted on top of each other. In addition to qualitative keel information, nine ice<br />

cores were extracted from several ridge slopes to record temperature and salinity profiles.<br />

The single most interesting feature studied was Ice Mountain, a large rubble pile (350' x 1100'),<br />

2.5 miles north of Narwhal Island in 59 feet of water (Figure 5). Probably formed by a single<br />

storm on St. Patrick's Day, a series of 3-4 pileups occurred with progressively less sail height.<br />

The maximum height measured was 72 feet. A formation of Ice Mountain is associated<br />

with a movement of more than 4000 feet which is estimated by studying seismic road<br />

separations immediately to the west, presented in more detail by Agerton (9).<br />

Another interesting, rubble pile occurred in 12' water depth inside the barrier islands within<br />

Prudhoe Bay, five miles south of Reindeer Island. The most dramatic ice feature inside the<br />

island during 1979, Rubble Pile (300' x I 000'), formed during early winter when the young ice<br />

sheet was only 4-8 inches thick. The moximum height of the pileup was 24 feet (Figure 6).<br />

Shear ridge was the third special first-year ice feature, extending from east of Narwhal to just<br />

northeast of Cross Island, approximately two miles offshore of the islands in 50-60 feet of<br />

water. Formed primarily by shearing action, the ice was ground up and refrozen, maintaining<br />

the shape of an almost vertical, straight wall 10 feet high.<br />

3. Aerial Photography<br />

Over 300 miles of stereo aerial photography was flown midway through the field investigation<br />

with the flight lines shown in Figure I. Flightline mosaics were developed and visually<br />

interpreted for aerial coverage of smooth ice, rubble, or ridges for each one-mile segment.<br />

Every third stereo pair of photographs was selected for centerline digitization and appeared to<br />

be sufficiently random for meaningful statistical analysis. Digitizing every third stereo pair<br />

permitted 50% coverage of each flightline due to 60% overlap between consecutive photo­<br />

graphs. Eighteen interesting special features located off the centerline were selected for<br />

additional digitization across or along them.<br />

To provide meaningful statistics-suitable criteria must be selected to scan the data and identify<br />

ridges and two-dimensional ice features. Previous criteria, including the "two-foot drop" used<br />

by Hibler (5) and others and the 50% Rayleigh criterion, have all overestimated the number of<br />

ridges. To determine ice volume estimates in pileup prediction on or force transmitted to,on<br />

offshore structure, it is necessary to differentiate between linear and two-dimensional ice<br />

features. Lowry and Wadhams (10) have developed such a criterion, but it assumes that all<br />

ridges have similar slopes which may not be the case.<br />

760


FIGURE 5 - ICE MOUNTAIN, LOOKING SOUTH<br />

FIGURE 6 - RUBBLE PILE, SHOWING THE WESTERN EXTREMITY<br />

LOOKING NORTH<br />

761


Part of the analysis in this study utilized a modified Rayleigh criterion (Criterion A), while<br />

another (Criterion B) was developed to distinguish between ridges and areal ice features.<br />

Criterion A. An ice feature starts if a data point is found to exceed a height of 3 feet.<br />

Subsequent data points are scanned to find the maximum height until a point is found to end the<br />

feature, indicating an elevation lower than one foot above smooth ice, or if it is in a trough<br />

which dropped by 80% of the maximum height. Criterion A is summarized in Figure 7.<br />

Criterion B. An ice feature ends only when a trough wider than 10 feet and lower than one foot<br />

high is encountered. Ridges were separated from other ice features by <strong>com</strong>paring the ratio of<br />

the height above the one-foot elevation and width of feature between one-foot elevations. An<br />

ice feature is assumed to be: I) rubble field if the maximum height-to-width ratio of the<br />

feature was less than 0.075; 2) rafted block if the ratio was greater than 0.5 and if the<br />

maximum height was less than 7 feet; 3) ridge, otherwise. These limiting conditions on<br />

maximum height-to-width ratios and maximum heights were selected from a parameter study<br />

verified by visually interpreting features from flight line V2. Criterion B is summarized in<br />

Figure 8.<br />

Histograms of ice features or ridges by height categories and population using both criteria<br />

were generated, and Table I shows the average population <strong>com</strong>parison between east-west and<br />

north-south f1ightlines. Additional analysis to <strong>com</strong>pare the ability of these criteria and others<br />

to predict extreme feature recurrences has been performed by Kreider and Thro (II).<br />

TABLE I -AVERAGE POPULATION COMPARISON BETWEEN LINES<br />

PERPENDICULAR AND PARALLEL TO THE COAST<br />

Perpendicular to Coost<br />

(Combining V I, V2, V3, and VS)<br />

Parallel to Coast<br />

(Combining H3, H4, HSB, HSA<br />

H6B, H6A, H7B, and H7A)<br />

4. Conclusions<br />

Population Per One-Fifth Mile<br />

Criterion A Criterion B Criterion B<br />

Ice Feature Ice Feature Ridges<br />

2.96 1.82 0.53<br />

1.95 1.52 0.43<br />

I) A new ice features identification criterion was developed to differentiate between linear<br />

ridges an two-dimensional ice features (e.g., rubble piles).<br />

2) A relationship has been established between sail height and keel depth for floating<br />

762<br />

ridges, while another relationship has been found to estimate the critical sail height<br />

representing the boundary between flooting and grounded ridges for a given water depth.


3) Rafted ice appeared to have greater areal coverage and thickness than previously<br />

considered.<br />

4) Most first-year ridges did not appear to be well-consolidated below six to eight feet<br />

below sea level wthin the keel.<br />

5) Large winter ice movements may occur in the outer perimeter of the landfast ice zone.<br />

5. Acknowledgments<br />

The authors thank the participating <strong>com</strong>panies (Arco, Chevron, Conoco, Exxon, Gulf, Mobil,<br />

Phillips, Shell) and their representatives for making the field program possible. A special<br />

acknowledgment goes to Gulf Research and Development Co. for their permission to present<br />

this paper.<br />

6 References<br />

I) Kovacs, A. (1971) "On Pressured Sea Ice," Sea Ice-<strong>Proceedings</strong> of an International<br />

Conference, Reykjavik, Iceland, p. 276-295.<br />

2) Rigby F. and A. Hanson (1976) "Evolution of a Large Arctic Pressure Ridge," AIDJEX<br />

Bulletin No. 34, Seattle, WA., p. 43-71.<br />

3) Weeks, W. F. and A. Kovacs (1970) "The Morphology and Physical Properties of Pressure<br />

Ridges: Barrow, Alaska - April 1969, "IAHR Ice Symposium <strong>Proceedings</strong>, Reykjavik,<br />

Iceland.<br />

4) Weeks, W. F., et al. (1971) "Pressure Ridge Characteristics in the Arctic Coastal<br />

Environment," PO AC-7 I <strong>Proceedings</strong>, Trondheim, Norway.<br />

5) Hibler, W. D. III, Mock. S. J., Tucker, W. B. III (1974) "Classification and Variation fo Sea<br />

Ice Ridging in the Western Arctic Basin", Journal of Geophysical Research, Vol. 79,<br />

pg. 2735-43.<br />

6) Wadhams, P. (1976) "Sea Ice Topography in the Beaufort Sea and its Effects on Oil<br />

Containment", AIDJEX Bulletin No. 33, pg. I-52.<br />

7) Tucker, W. B. III and Westhall, V. M. (1973) "Arctic Sea Ice Ridge Frequency Distribution<br />

Derived from Laser Profiles", AIDJEX Bulletin No. 21, pg. 171-ISO.<br />

S) Kan, T. K., et al. (1973) "Sonar Mapping of the Underside of Pack Ice", AIDJEX Bulletin<br />

No. 21, Seattle, WA., p. 155-169.<br />

9) Agerton, D. J. (19SI) "Major Nearshore Winter Ice Movements in the Alaskan Beaufort<br />

Sea", POAC-SI, Quebec City, Canada.<br />

10) Lowry, R. T., and Wadhams, P. (1979) "On the Statistical Distribution of Pressure Ridges<br />

in Sea Ice", Journal of Geophysical Research, Vol. S4, pg. 24S7-94.<br />

II) Kreider, J. R., and M. E. Thro (l9SI) "Statistical Techniques for the Analysis of Sea Ice<br />

Pressure Ridge Distributions", POAC-SI <strong>Proceedings</strong>, Quebec City, Canada.<br />

764


D.F. Dickins<br />

V.F. Wetzel<br />

MULTI-YEAR PRESSURE RIDGE STUDY<br />

QUEEN ELIZABETH ISLANDS<br />

DF Dickins Engineering<br />

Suncor Inc.<br />

Canada<br />

Canada<br />

ABSTRACT<br />

The study of multi-year pressure ridges within the Arctic Islands was conducted<br />

as Arctic Petroleum Operators Association Project 102. Sun Oil Company limited (now<br />

known as Suncor Inc.) was operator of the project and Norcor Engineering and Research<br />

limited was consultant. This project had as its primary objective the obtaining of<br />

fundamental data on multi-year pressure ridges in the Queen Elizabeth Islands of the<br />

Canadian Arctic. From a base camp on the ice in the Maclean Strait, 8 km west of<br />

Ellef Ringnes Island, the geometry and sail/keel depths of 12 free floating multiyear<br />

ridges was investigated. A rotating echo sounder transducer was used to<br />

determine the below water profile.<br />

A total of 20 ridge cross sections were obtained between May 11 and June 13, 1976.<br />

The mean keel/sail ratio of 5.6 ± 2.2 is larger and more variable than indicated<br />

from previous ridge studies conducted in more southern latitudes. A maximum keel<br />

of 37 meters was observed. Underwater keel profiles were characterized by a distinct<br />

asymmetry, with little correspondence between the location of maximum sail height<br />

and maximum keel depth along a particular ridge. Ice specific gravity, calculated<br />

from buoyancy considerations and a mean keel/sail ratio of 9.32, was 0.92. This<br />

corresponds almost exactly with quoted values for mUlti-year ice. The results of<br />

this study confirmed the abscence of significant void spaces in multi-year pressure<br />

ridges. This was in keeping with surface observations of fractured sails which<br />

clearly showed total consolidation of the original blocks. A trend toward higher<br />

sail/keel ratios with decreasing sail slope was evident, and this is postulated to be<br />

a manifestation of the ageing process over successive melt seasons.<br />

Further work will be required to better define the variability of keel/sail ratios<br />

and to determine the maximum keel depths that could be encountered in that region.<br />

This information is essential if we are to develop economic and environmentally<br />

safe designs for sub-sea production systems and pipelines in the Sverdrup Basin.<br />

765


INTRODUCTION<br />

Starting with the first exploratory hole drilled in the Arctic Islands in 1969,<br />

(Panarctic Drake Point K-67-A) to the Whitefish discovery of 1979, considerable<br />

interest has been shown in the development of specialized offshore drilling<br />

structures, pipelines, and suitable routes for ice strengthened tankers. Eventually,<br />

in the production phase, bottom founded well heads, terminal facilities and connector<br />

pipelines will have to be constructed. With safety as one of the primary environmental<br />

and operational concerns the probability of bottom scouring and impact from multiyear<br />

pressure ridge keels will be a major factor considered in the design of any<br />

facility or structure in this region.<br />

The study of multi-year pressure ridges was initiated by Sun Oil Company limited (now<br />

known as Suncor Inc.) through the Arctic Petroleum Operators Association as Project<br />

102. Canadian Superior, Gulf Canada Resources Inc. and Panarctic Oils were early<br />

participants with Phillips Petroleum and PetroCanada joining the study about a year<br />

later. Sun Oil Company limited was operator of the project with Norcor Engineering<br />

and Research limited as its consultant.<br />

Prior to this study, no investigation had been conducted concerning pressure ridge<br />

geometry in the High Arctic. The depth of ridge keels in areas of interest had to<br />

be estimated by applying a height to depth ratio of 1 to 3.2 (l, 2, 3). This ratio<br />

was determined from multi-year pressure ridge studies in the Beaufort Sea, and does<br />

not necessarily apply at more northern latitudes. The principal objective of this<br />

study was to determine the sail to keel ratio for various types of ice ridges <strong>com</strong>monly<br />

encountered in the Sverdrup Basin.<br />

766


The large degree of asymmetry typical of most of the ridges, meant that the maximum<br />

keel depth was often offset from the sail centre-line by 20 to 30 m. Figures 2 to 5<br />

show four different cross sectional profiles obtained along the length of one ridge.<br />

Profile 1 keel profile shows the excellent match between sonar data from two<br />

different sides. The asymmetry of the keel is important, as it indicates the uncertainty<br />

involved in estimating the entire ridge geometry from a half profile. The<br />

second profile, conducted 28 m further west, incorporated the ridge high point of<br />

6.4 m. Here, the deepest keel of the programme was observed at 40 m. There was<br />

still a faint sonar return at this depth, but lack of more sonar rods prevented a<br />

transducer mounting lower than 30 m. This extremely deep section is probably a<br />

localized feature, because on the matching north profile, the return was lost at 35 m.<br />

The third profile was undertaken to prove whether the keel depth was at all proportional<br />

to the sail at a particular point. A position was chosen off the main point<br />

of the ridge where the sail was very wide (60 m) and low (3 m). No significant keel<br />

was expected here, but the sonar record showed a keel more massive than Profile I,<br />

extending 32 m into the water. With the transducer perpendicular to the ridge, the<br />

keel bottom was below the limit of sonar again, i.e., > 40 m. The nearly vertical<br />

face indicates that part of the original keel was missing.<br />

Profile 4 was conducted still further to the East where the ridge was extremely broad<br />

and only had a 2 m sail projecting above the general elevation of the multi-year floe.<br />

As on Profiles 1 and 2, this keel was asymmetric, with most of the mass to the north<br />

of centre. The keel is extremely irregular and could not have experienced significant<br />

ablation.<br />

769


etween 30 and 50 m depth (8). Other oceanographic measurements, taken in Parry<br />

Channel, show a similar temperature stratification at the same depth range.<br />

Consequently, the majority of ridges in the study area have their keels within<br />

water below its freezing point during much of the winter period. The open water<br />

season in the summer, north of 78°N is very short and often non-existent. Compared<br />

to ridges in more southern areas, there seems to be much less opportunity for keel<br />

depletion in the Queen Elizabeth Islands area.<br />

For impact purposes, the total mass of a typical ridge is of interest. The three<br />

areas of the Ridge 13 profile were averaged. When applied to a 200 m long ridge,<br />

the mass would be 340 (10)6 kg • A corresponding smooth piece of mUlti-year ice 6 m<br />

thick x 200 m x 100 m (approximate ridge keel width) would weigh 140 (10)6 kg • The<br />

presence of the ridge has effectively increased the ice mass by a factor of 2.4.<br />

CONCLUSIONS<br />

The data presented in this paper is an extraction from a report containing the entire<br />

body of current knowledge about multi-year pressure ridge geometry in the Queen<br />

Elizabeth Islands of the Canadian Arctic.<br />

The mean sail/keel ratio found in this study was 1 to 5.6 ± 2.2 (population of 17).<br />

This value is considerably larger and more variable than indicated by previous ridge<br />

studies at more southern latitudes. The mean total ice thickness of the ridges<br />

studies here was 25.3 m with a maximum thickness of 46 m. It is theorized that the<br />

different aging process of ridges in the islands area, due to cold air and water<br />

temperatures over most of the year, contributes to the large sail to keel ratios<br />

found here.<br />

The ridges studies were generally in hydrostatic equilibrium and underwater profiles<br />

were characterized by a distinct asymmetry. A trend toward higher sail/keel ratios<br />

with decreasing sail side slope angle was evident, and this is postulated to be a<br />

result of the aging process over successive melt seasons.<br />

Due to the high sail/keel ratios found in this study, considerable re-thinking may<br />

be necessary to accurately present the operational and design constraints imposed<br />

by ridging in the Arctic Islands area. With similar ratios even small ridges of 3 m<br />

sail height, not previously considered a severe obstacle, may contain over 20 m of ice.<br />

774


REFERENCES<br />

1. KOVACS, A., 1972: On Pressured Sea Ice, Sea Ice Conference <strong>Proceedings</strong>,<br />

Reykjavik, Iceland.<br />

2. KOVACS, A., 1973: Structure of a Multi-Year Pressure Ridge, ARCTIC, Volume 26,<br />

No.1.<br />

3. KOVACS, A., DICKINS, D.F., WRIGHT, B., 1975: An Investigation of Multi-Year<br />

Pressure Ridges and Shore Pile-Ups. A.P.O.A. Project 89 by NORCOR Engineering<br />

for GULF CANADA LTD.<br />

4. WETZEL, V.F., 1971-75: Statistical Study of Late Winter Ice Thickness<br />

Distribution in the Arctic Islands, A.P.O.A. Project 96 - proprietary data.<br />

5. ANDERSON, D.L., 1960: The Physical Constants of Sea Ice, Research, Volume 13,<br />

pp. 310-18.<br />

6. DOAKE, C.S.M., 1976: Thermodynamics of the Interaction Between Ice Shelves and<br />

the Sea, Polar Record, Volume 18, No. 112.<br />

7. SVERDRUP, H.U., no date: Arctic Sea Ice, Encyclopedia Arctica, Vol. 7,<br />

unpublished manuscript.<br />

8. FUJINO, K., LEWIS, E.L., PERKIN, R.G., 1974: The Freezing Point of Seawater at<br />

Pressures up to 100 Bars, Journal of Geophysical Research, Volume 79, No.<br />

12.<br />

9. DICKINS, D.F., 1976: Multi-Year Pressure Ridge Study, Queen Elizabeth Islands,<br />

A.P.O.A. Project 102 report by Norcor Engineering and Research Ltd. to<br />

Sun Oil Company Ltd.<br />

775


L. Wolfson,<br />

Senior Drilling Engineer<br />

-Ice Studies Aid in the Successful COmpletion<br />

of the Norton Sound C.O.S.T. Well-<br />

ARCO Oil and Gas Company<br />

W. M. Evans,<br />

Senior Staff Drilling Engineer ARCO Oil and Gas Company<br />

Abstract<br />

U.S.A.<br />

U.S.A.<br />

ARCO Oil and Gas Company was the operator of a C.O.S.T. Well drilled in Norton Sound<br />

during the open water season of 1980. Satellite imagery were used to document<br />

historical ice breakup and ice freezeup periods. Meteorological data were used to<br />

determine causal effects during these periods and forecast models were developed to<br />

predict ice breakup and ice freezeup. Historical and real time ice and meteorological<br />

data were used to forecast and advise of favorable operating conditions for 1980. Other<br />

studies and activities which led to the successful <strong>com</strong>pletion of the well included:<br />

(1) determination of extreme and normal meteorological and oceanographic conditions in<br />

the open water season; (2) establishment of supply facilities available and scheme for<br />

supply support for the well; (3) sea floor condition; (4) establishing a well prognosis<br />

which included well design, drilling time and cost estimate; and (5) the rig<br />

mobilization and demobilization plan.<br />

Introduction<br />

In August 1978, ARCO Oil and Gas Company, as operator for a group of petroleum industry<br />

participants, initiated plans to drill a Continental Offshore Stratigraphic Test<br />

(C.O.S. T.) Well in the Norton Sound Basin, Offshore Alaska, Figure 1. To facil itate the<br />

planning and operations of the C.O.S.T. well, several studies and various activities<br />

were initiated.<br />

776


A jack-up drilling vessel, the Dan Prince, was chosen to drill the C.O.S.T. well. The<br />

Dan Prince was capable of operating only in an ice-free environment. To determine if<br />

sufficient time was available to <strong>com</strong>plete the well in one season, a study to document<br />

the historical ice-free season was conducted. To take maximum advantage of the open<br />

water season, forecast models were developed to predict the onset of ice breakup and ice<br />

freezeup.<br />

Additional studies and activities were initiated to aid in logistical and operational<br />

planning and the permitting process. A site specific seismic survey was conducted to<br />

determine if potential shallow drilling hazards were present and also prepare the<br />

drilling prognosis. Sea floor soil conditions were evaluated to determine if the soils<br />

would provide an adequate foundation for the rig. A tow efficiency study was conducted<br />

to assist in planning mobilization and demobilization of the rig. Extreme and normal<br />

meteorological and oceanographic conditions were established to evaluate the suitability<br />

of the rig for the environment in this area and to evaluate supply schemes.<br />

Logistical requirements were determined to assure timely supply of material and<br />

movement of personnel.<br />

Ice Breakup and Ice Freezeup Documentation<br />

Since the drilling vessel was able to operate only in an ice-free environment, the<br />

vessel's owners, insurance underwriters and regulatory agencies needed assurance that<br />

operations would be conducted only during the ice-free season. This dictated the need<br />

to establish adequate forecasting of the ice breakup and ice freezeup periods.<br />

Historical ice breakup and ice freezeup periods and the associated meteorological data<br />

were documented and used for developing ice forecasting models for this area.<br />

Ice conditions in the Northern Bering Sea and Norton Sound are quite varied with respect<br />

to ice features, concentration, and movement on a spatial and temporal basis. The area<br />

of the C.O.S.T. well site can be ice-free at certain times, but ice from other areas can<br />

be driven into the area by changing meteorological conditions. To establish the ice<br />

breakup and ice freezeup periods in areas in the vicinity and at the C.O.S.T. well site,<br />

four areas which are illustrated in Figure 2 were studied: Area 1, Northern Bering Sea<br />

(St. Lawrence Island to the Bering Strait); Area 2, Norton Sound (East of St. Lawrence<br />

Island); Area 3, St. Lawrence (South of a line from St. Lawrence Island to the Yukon<br />

Delta); and Area 4, Bays and Inlets (Go1vin Bay and Norton Bay).[1]<br />

777


if it can be used for long range forecasts to a certain degree of accuracy. In<br />

mid-March, the first ice breakup forecast was issued using the pressure data for the<br />

first two weeks of March. The first freeze up forecast was issued on the first of<br />

September and pressure data for the last two weeks of August were used in the forecast<br />

model. Subsequent forecasts were made at two week intervals and the pressure data used<br />

were the data used for the previous forecasts plus the pressure data recorded during the<br />

most recent two week period. For the 1972-1978 data base, the models developed<br />

accurately predicted the ice breakup and ice freeze up dates established in the study.<br />

It was recognized that such a small data base (seven years) for relatively average<br />

weather conditions during the seven years might not be representative for years that<br />

deviate considerably from average.<br />

To test the accuracy of the forecast models, a study of the performance of the models<br />

for predicting ice breakup and ice freezeup for the 1979 season was initiated.[2,3] As<br />

noted previously, ice conditions during 1979 were much less severe. The ice breakup<br />

model performed poorly; the ice freezeup model was more accurate than the ice breakup<br />

model but did not approach the degree of accuracy established in the 1972-1978 study.<br />

This 1979 study illustrated the need of the forecasters to use the predictive models<br />

only as a guide. The forecasts from the models would have to be modified by the forecast<br />

personnel relying on their personal experience using all available meteorological and<br />

satellite data on a real time basis. These modified forecasts are referred to as<br />

subjective forecasts.<br />

Both model and subjective semi-monthly forecasts were issued beginning on March 19,<br />

1980 for support of the 1980 drilling program.[4] The subjective forecasts for ice free<br />

condtions in Area 2 (which includes the C.O.S.T. well site) were usually within one week<br />

of the actual ice free date. The model forecasts for ice free conditions varied from<br />

one and one-half weeks to three weeks.<br />

Semi-monthly forecasts for ice freezeup were initiated in late August. As with the ice<br />

breakup forecasts, the model forecasts varied with the most accurate forecast being the<br />

first forecast on September 1 (within 4 days) which had the least data base. The error<br />

in subsequent forecasts increased. Subjective forecasts were very accurate and the<br />

first forecast issued in late August was within 2 days of the inception of freeze up in<br />

Norton Sound. Subsequent subjective forecasts were not in error.<br />

779


Utilization of Ice Data for Actual Operations<br />

What is important is not only what the forecasts were and the accuracy of the forecasts,<br />

but how the operator used these data to mobilize the jack-up rig to Norton Sound. The<br />

first subjective forecast indicated that open water conditions would be expected in<br />

Area 2 by June 10 and the area should be ice free by June 28. 8y mid-March, 1980, ARCO<br />

anticipated bein9 able to mobilize the jack-up rig in time to start operations by<br />

mid-June. A meeting was held with the contractor and marine surveyor in mid-March to<br />

review all prior work that had taken place in regard to establishing the ice free season<br />

variations and how the forecasts of ice breakup for 1980 would be utilized.<br />

Subsequently, real time satellite data and ice charts prepared from the satellite data<br />

were reviewed on a systematic basis by operator and contractor personnel. Cloud cover<br />

prevented monitoring the C.O.S.T. well site from satellite photos shortly before the<br />

rig arrived at the site. During this time, overflights were made of the area to assure<br />

that the site was ice free and to monitor the movement of ice in Area 1 (Northern Bering<br />

Sea). The drilling rig was secured on location in Norton Sound on June 13, 1980 without<br />

encountering sea ice.<br />

Ice freezeup forecasts which were initiated in late August indicated that ice would<br />

start to form in Norton Sound in late October. Ice was first observed on October 27,<br />

1980. Operations were <strong>com</strong>pleted on September 29, 1980, thus the well activities were<br />

finished prior to ice forming in the area.<br />

Other Studies and Activities<br />

A. Site Specific Seismic Surveys<br />

Sea floor, shallow seismic and deep seismic surveys were made in the summer of 1979.<br />

Sea floor surveys were used to determine the possibility of encountering sea floor<br />

hazards such as boulders and gas seeps. Shallow seismic surveys were used to determine<br />

(1) the thickness and continuity of shallow formations, (2) if faults were present and,<br />

(3) if shallow high pressure gas zones were present. Data from the deep seismic surveys<br />

were used to establish (1) if the formations to be penetrated would be expected to be<br />

normally or abnormally pressured, (2) the casing program, and (3) the expected<br />

formation drillability. The formation drillability data were used in conjunction with<br />

the proposed drilling, well logging and coring program, to establish the expected total<br />

well time. The expected well time was determined to be 126 days and the actual total<br />

well time was 107 days.<br />

780


B. Geotechnical Data[S]<br />

In 1979, cores were taken at the sea floor to correlate the shallow soil conditions with<br />

the shallow seismic survey. These data were used to establish the suitability of the<br />

soil conditions for supporting the legs of the jack-up rig. In establishing the<br />

suitability for jack-up rigs, soils need to be of sufficient strength to support the<br />

weight of the legs but allow for some leg penetration to prevent the drilling rig from<br />

sliding if subjected to high horizontal loading from environmental conditions. The<br />

corehole data were also used to plan the setting of the shallow casing string. This<br />

casing string acts as a foundation for the other casing strings.<br />

C. Tow Efficiency Studies[6,7]<br />

Previous experience offshore Alaska indicated that very high localized wind conditions<br />

could develop quickly offshore, especially in close proximity to mountain ranges. The<br />

jack-up rig was located in the Lower Cook Inlet and was to be towed south along the<br />

Alaska Peninsula through Unimak Pass enroute to Norton Sound (Figure 4). A tow<br />

efficiency study was initiated to determine the expected time to mobilize the vessel to<br />

Norton Sound. Six different starting times were used beginning with the first week in<br />

May and the first day of five successive weeks. Weather data used were for the years<br />

1974 through 1979. The study showed very little variation in expected tow times because<br />

the weather conditions were not expected to be severe.<br />

One does not know for sure what weather conditions might be encountered during tow<br />

operations. The tow efficiency study also documented "safe havens" along the route<br />

where the rig could be diverted in the event of a storm of such magnitude which would<br />

dictate going to a sheltered area. The results of the tow efficiency study and the<br />

identified "safe havens" were reviewed with operator and contractor personnel involved<br />

with the operations prior to the rig leaving the Lower Cook Inlet. Marine forecasts<br />

were disseminated to the tug and barge personnel on a routine basis during the tow. No<br />

adverse weather conditions were experienced during the tow to Norton Sound. The<br />

expected tow time was 13 days but the actual tow time was 10.4 days. A tow speed of 4.7<br />

knots in calm seas was used in the simulation. The approximate average tow speed was<br />

S.2 knots.<br />

On <strong>com</strong>pletion of drilling the C.O.S.T. well, there was the possibility of moving the rig<br />

back to the Lower Cook Inlet. A similar tow efficiency study was made for moving the<br />

rig to the Lower Cook Inlet using eight different starting dates from September 14<br />

781


through November 12. The study indicated that adverse weather conditions could be<br />

expected and possible delays enroute could result. The rig was subsequently<br />

demobilized to Dutch Harbor in the Aleutian Islands and then released by ARCO. Tow time<br />

from Norton Sound to Dutch Harbor was approximately 5.3 days.<br />

While the rig did not have to be diverted to "safe havens", the above studies were<br />

instrumental in informing the contractor and marine surveyor of the difficulty that<br />

might be experienced in such tow operations offshore Alaska. For subsequent equipment<br />

mobilization in these areas, ARCO will initiate similar studies.<br />

D. Extreme and Normal Meteorological and Oceanographic Conditions[8,9]<br />

Extreme meteorological and oceanographic conditions were determined for the open water<br />

months of June through October to evaluate the suitability of the jack-up rig for use<br />

at the particular water depth, in the given soil conditions. Marine surveyors make<br />

calculations to assure (1) the rig will not fail by overturning, (2) environmental<br />

forces will not exceed the yield strength of the legs, (3) there is sufficient soil<br />

strength to prevent the lower support (cans or mat) from penetratin9 the soil further<br />

during environmental loading and, (4) the rig will not slide along the sea floor under<br />

certain environmental conditions. The calculations are made for certain calculated le9<br />

and can (or mat) penetrations of the sea floor and for a given deck or barge clearance.<br />

In areas where oil activities have been carried out for a number of years, the 50 year<br />

extreme event storm is sometimes used. In frontier areas, some marine surveyors use 100<br />

year extreme event storm data.<br />

An extreme event analysis was made by hindcasting severe storms which affected the<br />

Norton Sound area in the time frame of 1955 through 1978 (24 years). Extreme events for<br />

various return intervals were determined. The 100 year return interval extreme event<br />

was used to determine the suitability of the jack-up rig Dan Prince for operations in<br />

Norton Sound.[10]<br />

A hindcast of normal wind and wave conditions was made to determine the expected<br />

environmental operating conditions in the open water months of June through October.<br />

The hindcast involved determining wind and wave conditions for consecutive 6-hour<br />

periods during a particular year. A total of three years were hindcast. These data<br />

were used to determine (1) the expected environmental conditions when the rig arrived<br />

on location, (2) if supply of the rig would be difficult due to expected environmental<br />

conditions, and (3) the expected environmental conditions when operations would have to<br />

be terminated.<br />

782


Even though the environmental conditions for Norton Sound are milder than most areas<br />

offshore Alaska, the hindcast study did provide data to evaluate potential problems<br />

during raising the barge at the start of operations and lowering the barge at the end<br />

of operations. Raising and lowering the barge of a jack-up rig requires mild sea states<br />

in order to prevent high impact loads of the legs with the sea floor. The hindcast data<br />

were also used to determine the feasibility of using a supply barge moored close to the<br />

jack-up rig.<br />

E. Logistics<br />

Logistical operations involved the use of two helicopters, a helicopter base and<br />

expeditor office at the Nome airfield, a 1ightering tug and barge for transporting<br />

fresh water from Nome to supply boats offshore, 2 supply boats and a large supply barge.<br />

Water depth at Nome (approximately 6 feet) prevented the use of regular draft supply<br />

boats (16-19 feet draft) at the Nome harbor. The 1i9htering barge was approximately<br />

120 feet long, 40 feet wide, 8 feet high and had a draft of approximately 3 feet.<br />

A supply barge (400 ft. long, 76 ft. wide, 20 ft. high) was the major supply base for<br />

drilling materials and was moored approximately three miles from the rig. The barge was<br />

equipped in Seattle and towed to the well site. The barge was equipped with living<br />

quarters, a sewage treatment plant, fresh water and fuel oil tanks, refrigeration<br />

equipment, generator, crane, forklift, bulk storage tanks for mud and cement, casing<br />

and miscellaneous drilling equipment.<br />

Mooring requirements of a barge in an open seaway are greater than mooring requirements<br />

of a barge in protected waters. A mooring study was made and a single point mooring<br />

system was designed for the barge.[II] The mooring system was designed to withstand a<br />

25-year return interval storm. In the event the mooring system failed, an emergency<br />

anchor was attached to a wire1ine and winch system which would allow for rapid<br />

deployment of the emergency anchor. The emergency anchor was incorporated into the<br />

mooring system design to prevent the barge from drifting free in the event the primary<br />

system failed. Also, a <strong>com</strong>plete backup mooring system was available in the event the<br />

primary mooring system was lost.<br />

These supply arrangements allowed for activities to be conducted on the rig without<br />

delays due to inadequate supplies.<br />

783


Conclusions<br />

The C.O.S.T. well was successfully <strong>com</strong>pleted on September 29, 1980. The utilization of<br />

the ice studies, site survey and geotechnical data, meteorological and oceanographic<br />

data and the supply scheme allowed for drilling personnel to utilize the drilling rig<br />

and drilling scheme with a minimum of problems.<br />

Acknowledgements<br />

The authors wish to acknowledge the contributions of Kwang U. Park, ARCO Oil and Gas<br />

Company and Gary M. Wohl, Oceanographic Services, Inc.<br />

REFERENCES<br />

1. Oceanographic Services, Inc: "Freezeup and Breakup Forecasting, Norton Sound",<br />

July, 1979<br />

2. Oceanographic Services, Inc.: "Norton Sound Ice Forecasting, 1979 Breakup and<br />

Freezeup Predictions and Documentation", March, 1980<br />

3. Oceanographic Services, Inc.: "Norton Sound Ice Forecasting, 1979 Breakup and<br />

Freezeup Statistical Verification and Supplement", March, 1980<br />

4. Oceanographic Services, Inc.: "Norton Sound Ice Forecasting, 1980 Breakup and<br />

Freezeup Predictions and Documentation (Preliminary Report)", December, 1980<br />

5. Woodward-Clyde Consultants: "Geotechnical Investigation Program and Siting<br />

Study for Jack-up Rig Footing Behavior in Norton Sound, Offshore Alaska",<br />

October, 1979<br />

6. Oceanroutes, Inc.: "Tug/Tow Simulation; Kachemak Bay, Alaska To Norton Sound<br />

Alaska," April,1980<br />

7. Oceanroutes, Inc.: "Tug/Tow Simulation; Norton Sound, Alaska to Kachemak Bay,<br />

Alaska," May, 1980<br />

8. Oceanographic Services, Inc.: "Environmental Study Norton Sound," December,<br />

1978<br />

9. Oceanographic Services, Inc.; "Extreme Event Analysis, Norton Sound C.O.S. T.<br />

Well Site," March, 1980<br />

10. Noble Denton & Associates Inc.: "Self-Elevating Offshore Drilling Platform "Dan<br />

Prince", Re<strong>com</strong>mendations for Operations Offshore Alaska in Areas of the Cook<br />

In 1 et and Norton Sound," February, 1980<br />

11. J. Ray McDermott & Co. Inc.: "Mooring System Design for Supply/Storage Barge,<br />

Norton Sound, Alaska," April 21, 1980<br />

784


.....<br />

'"<br />

BERING STRAIT<br />

NORTHERN<br />

BERING SEA<br />

AREA<br />

3<br />

ST. LAWRENCE AREA<br />

2<br />

NORTON SOUND<br />

AREA<br />

FIGURE 2. - ICE FORECAST AREAS<br />

NORTON SOUND


LEGEND<br />

Ice Concentration in Octas (8ths)<br />

IF: Ice Free<br />

OW: Open Water<br />

F: First Year Ice<br />

V: Young Ice<br />

N: Nilas<br />

O.W.<br />

BERING STRAIT<br />

6-7<br />

l-:h',-SF<br />

7-8<br />

IV, 6-7<br />

6-7<br />

IV, S--6F<br />

fiGURE 3 .. TYPICAL ICE CHART


788<br />

FIGURE 4 . TOW SIMULATION ROUTING


J. R. Kreider<br />

M. E. Thro<br />

STATISTICAL TECHNIQUES FOR THE ANALYSIS OF<br />

SEA ICE PRESSURE RIDGE DISTRIBUTIONS<br />

Shell Development Company<br />

Shell Development Company<br />

U.S.A.<br />

U.S.A.<br />

A B S T R ACT<br />

Techniques to obtain pressure ridge statistics are evaluated using stereoaerial<br />

photography data from the nearshore Alaskan Beaufort Sea. Four ridge definitions<br />

are <strong>com</strong>pared. Similar probability distributions for ridge height result, but<br />

significant differences for the number of ridges per mile occur. A first-order negative<br />

exponential distribution is found to fit the sail height data. A Type I Extreme<br />

Value Distribution for maximum ridge heights is proposed and found to fit existing<br />

data.<br />

I N T ROD U C T ION<br />

In April 1979, Gulf Research and Development Company, as administrator for a<br />

joint industry project supported by eight oil <strong>com</strong>panies, acquired stereo-aerial photography<br />

in the nearshore Alaskan Beaufort Sea [1]. Centerline profiles of these photographs<br />

were digitized, from which statistics on first-year pressure ridges and rubble<br />

were <strong>com</strong>puted. These statistics can be used to establish extreme ridge thicknesses<br />

for structural design and to plan surface logistics for winter operations.<br />

Determining pressure ridge statistics requires a ridge definition to identify<br />

independent ridges. Several different ridge definitions and probability distributions<br />

for ridge height have been used previously. The objective of this paper is to<br />

assess the effect of analysis technique on the calculated ridge statistics.<br />

A N A L Y SIS T E C H N I QUE S<br />

D a t a S e 1 e c t ion<br />

Figure 1 shows the flightlines in the nearshore Alaskan Beaufort Sea near<br />

Prudhoe Bay. Centerline profiles are digitized from one-mile stereo models of the<br />

photographs [1]. Fifty percent of the total line length is digitized by alternately<br />

digitizing and skipping one-mile segments.<br />

789


z<br />

0<br />

I-<br />

U<br />

z<br />

::><br />

l.L<br />

>-<br />

I-<br />

(f)<br />

Z<br />

W<br />

0<br />

>-<br />

I- 10-2<br />

-.J<br />

!D<br />

([<br />

!D<br />

0<br />

a:<br />

D..<br />

10-3<br />

3<br />

o SOX RRYLE I GH<br />

D 2-FT DROP<br />

+ RRYLE I GH \I ITH<br />

7" SLOPE<br />

)( \-FT TROUGH<br />

ELEVATION<br />

)(<br />

W x<br />

6- x<br />

S = 0.625 FT -\---<br />

5 7 9 11<br />

RIDGE HEIGHT<br />

h (FT)<br />

Figure 2a<br />

13 0 100 200<br />

R lOGE HE I GHT<br />

SQUARED h 2 (FT2)<br />

Figure 2b<br />

given in Table 1, indicate that both distributions should be accepted for all defini­<br />

tions at the 95 percent significance level.<br />

For the nearshore zone, the data fit equation (3) up to a ridge height of<br />

approximately ten feet. At larger ridge heights, a few single ridges distort the<br />

distribution. This distortion may be due to a relatively small data sample or to the<br />

fact that the larger sail heights are associated with grounded ridges, which will have<br />

proportionately larger sails for the same volume of deformed ice.<br />

Previous investigators have found both definitions to apply to sail and keel<br />

data. Hibler et a1. [10] finds that equation (1) fits both keel depth and sail height<br />

distributions in the Central Arctic Basin. Wadhams [5] finds that equation (1) fits<br />

keel data and that equation (3) fits sail data for the same area in the Arctic Ocean<br />

northeast of Greenland. Wadhams [4] and Tucker et a1. [6] find that equation (3) best<br />

fits sail data in the coastal Beaufort Sea. Our results indicate that equation (3)<br />

provides a better fit, although both distributions pass a Chi-squared test.<br />

ENG I NEE R I N G CON SID ERA T ION S<br />

Rid g e S ail S tat i s tic s<br />

An important consideration for engineering applications is the probability<br />

of exceedance, or nonexceedance, for a given ridge height. If the probability dis­<br />

tributions for ridge occurrence and ridge height are known, the probability of non­<br />

exceedance can be calculated as follows:<br />

793


5. Wadhams, P. 1977. "A Comparison of Sonar and Laser Profiles along Corresponding<br />

Tracks in the Arctic Ocean". AIDJEX/ICSI Symposium. Sea Ice Processes and<br />

Models, University of Washington Press, 283-299.<br />

6. Tucker, W. B. III, W. F. Weeks, and M. D. Frank. 1979. Sea Ice Ridging over the<br />

Alaskan Continental Shelf. CRREL Report 79-8.<br />

7. Tucker, W. B. III and V. H. Westhall. 1973. Arctic Sea Ice Ridge Frequency<br />

Distribution Derived from Laser Profiles. AIDJEX Bulletin, 21, 171-180.<br />

8. Hibler, W. D. III, S. J. Mock, and W. B. Tucker III. 1974. Classification and<br />

Variation of Sea Ice Ridging in the Western Arctic Basin, Journal of Geophysical<br />

Research, 79, 18, 2735-2743.<br />

9. Lowry, R. T. and P. Wadhams. 1979. On the Statistical Distribution of Pressure<br />

Ridges in Sea Ice. Journal of Geophysical Research, 84 (C5), 2487-2494.<br />

10. Hibler, W. D. III, W. F. Weeks, and S. J. Mock. 1972. Statistical Aspects of<br />

Sea-Ice Ridge Distributions, Journal of Geophysical Research, 77 (30), 5954-<br />

5970.<br />

11. Hibler, W. D. III. 1975. Statistical Variations in Arctic Sea Ice Ridging and<br />

Deformation Rates. <strong>Proceedings</strong> of the Ice Technology Symposium, Montreal, 9-11,<br />

April, pp. Jl-J19, Society of Naval Architects and Marine Engineers, New York.<br />

12. Keinonen, A. 1976. The Shape and Size of Ridges in the Baltic According to<br />

Measurements and Calculations, Winter Navigation Research Board Report No. 17,<br />

Helsinki, Finland.<br />

13. Ackley, S. F., W. D. Hibler III, F. K. Kugzruk, A. Kovaks, and W. F. Weeks.<br />

1974. Thickness and Roughness Variations of Arctic Multiyear Sea Ice. AIDJEX<br />

Bulletin, 25, 75-96.<br />

798


Gordon F.N. Cox·<br />

W.S. Delm<br />

Abstract<br />

SlM1ER ICE CONDITIOOS IN 1HE<br />

PRUDHOE BAY AREA, 1953-75<br />

US Army Cold Regions Research<br />

and Engineering Laboratory<br />

Sea Ice Consultants<br />

A detailed knowledge of the summer ice conditions is required for planning offshore<br />

petroleum operations in the Prudhoe Bay area. Statistics on breakup and freezeup<br />

dates and the number of open water days are needed to plan and assess the feasibility<br />

of fill island and causeway construction, pipeline laying, platform installation,<br />

and other summer activities such as seismic vessel operations. Breakup and freezeup<br />

data are also needed to schedule winter operations.<br />

Long-term, site-specific statistics on the summer ice conditions in the Harrison<br />

Bay - Camden Bay area are presented in probalistic terms. The statistics are based<br />

on twenty-three years of ice observations acquired by <strong>com</strong>mercial ships and icebreakers,<br />

ice reconnaissance flights, and various satellites. Data is given on<br />

breakup and freezeup dates, the first occurrence of open water, and the number of<br />

continuous and total open water days. The impact of the summer ice conditions on<br />

petroleum activities in the study area are also briefly discussed.<br />

*Most of this work was prepared while employed by Amoco Production Company,<br />

Research Center, Tulsa, OK. 799<br />

USA<br />

USA


1. Introduction<br />

A historical perspective of the summer ice conditions along the north coast of Alaska<br />

is needed to evaluate the feasibility and cost of operating in that area during the<br />

summer. Long-term statistics on river break-up and overflow, sea ice break-up,<br />

floe size, ice concentrations, and pack ice invasions are needed for planning: barge<br />

and ship navigation; seismic vessel operations; platform towing and installation;<br />

gravel island and causeway construction; and pipe laying operations. Break-up and<br />

freeze-up data are also needed to schedule winter operations.<br />

In 1976 the Research Department of Amoco Production Company and Sea Ice Consultants<br />

performed an in-depth analysis of the summer ice conditions along the central,<br />

third portion of the north Alaskan coast. The study area extended from Harrison<br />

Bay to Camden Bay, out to the continental slope (Figure 1). Twenty-three years<br />

(1953 to 1975) of summer ice conditions data were <strong>com</strong>piled for the study area from<br />

all available data sources and site-specific ice statistics were generated for<br />

twenty sites inside the 20-meter isobath. The results were later made available<br />

to interested parties as Alaskan Oil and Gas Association, Project 35.<br />

This paper discusses the data and methods used in the Amoco study. Statistics on<br />

sea ice break-up and freeze-up dates, the first occurrence of open-water, and the<br />

number of continuous and total open-water days are presented for each of the twenty<br />

sites. The impact of the summer ice conditions on petroleum activities in the<br />

study area is also discussed.<br />

2. Previous Work<br />

Regional information on the summer ice conditions off the Alaskan north coast has<br />

been summarized by Potocsky (1). Potocsky used the Naval Oceanographic Office<br />

Annual Ice Reports to calculate the mean, median, ranges of the l5-day mean, and<br />

extreme southern and northern positions of the pack ice edge. This was done for<br />

semi-monthly periods at 50 intervals of longitude along the coast. Landsat<br />

satellite imagery have also been used by several investigators to determine the<br />

extent and variation of the open-water season off the central Alaskan coast (2,3).<br />

While these studies provide an overview of the ice conditions, the results do not<br />

lend themselves to detailed site-specific analyses. The ice charts in the Navy<br />

annual ice reports which were used by Potocsky only provide a regional description<br />

of the ice conditions. Even though the resolution of Landsat emagery is more than<br />

adequate, the Landsat data base is limited by few years of operation and the low<br />

frequency of passes over an area of interest. Often the ice is obscured on Landsat<br />

imagery by cloud cover.<br />

3. Data Sources<br />

In order to obtain an adequate data base for detailed, site-specific analyses of<br />

the summer ice conditions, it was necessary to consider all available ice data<br />

sources. The data sources used in this investigation are listed below:<br />

800<br />

1. Original U.S. and Canadian ice observer flight logs and messages<br />

2. U.S. and Canadian annual ice reconnaissance reports<br />

3. Commerical and government ship reports<br />

4. Satellite imagery<br />

a. NOAA and ESSA SR (Scanning Radiometer)


00<br />

o<br />

....<br />

71°30'<br />

71°<br />

70°30'<br />

70 0<br />

N<br />

Beaufor Sea<br />

Figure 1: Study Area and Twenty Sites Chosen for Statistical Analyses.<br />

144°W<br />

71° 30'


00<br />

o<br />

'"<br />

71"30'<br />

70"30'<br />

70"N<br />

o 7<br />

T5T<br />

Beaufor Sea<br />

4MY<br />

3FT<br />

PD3<br />

Figure 2: Example of an Ice Chart Used in This Study, WM) Nomenclature<br />

is used to describe the ice conditions (7),<br />

Undercast<br />

144"W<br />

144"W<br />

71"30'


Because of a lack of observational data for some periods, approximately 5% of the<br />

ice charts were prepared by hindcasting the ice conditions. Hindcasts were performed<br />

using surface pressure charts, mean wind data, and temperature records. Previous and<br />

subsequent ice conditions were taken into consideration.<br />

After the ice charts were prepared, data on average weekly ice concentration, ice<br />

type, and floe size were obtained for 20 sites in the study area (Figure 1) and<br />

presented in time series form for subsequent statistical analyses. One week was<br />

defined as a six-day period so that the results could be <strong>com</strong>pared to the U.S.<br />

Navy annual ice reports. The first period or week represents the first six days in<br />

June, and so on (Table 1).<br />

Period Time<br />

1 June 1 to 6<br />

2 7 12<br />

3 13 18<br />

4 19 24<br />

5 25 30<br />

6 July 1 to 6<br />

7 7 12<br />

8 13 18<br />

9 19 24<br />

10 25 30<br />

11 31 5<br />

12 August 6 to 11<br />

13 12 17<br />

14 18 23<br />

15 24 29<br />

16 30 4<br />

17 September 5 to 10<br />

18 11 16<br />

19 17 22<br />

20 23 28<br />

21 29 4<br />

22 October 5 to 10<br />

23 11 16<br />

24 17 22<br />

25 23 28<br />

Table 1: Time during summer season corresponding to a given period.<br />

5. Statistical Ana1lses<br />

Various types of statistical analyses were performed as the required ice<br />

statistic depends on the nature of the operation of interest. Due to space<br />

limitations on the length of this paper, only a few general statistics will be<br />

presented here. These include statistics on break-up, the first occurrence of<br />

open-water, freeze-up, and the number of continuous and total open-water days<br />

804


during the summer. Other useful statistics for a 10-meter water depth site inside<br />

the barrier islands near Prudhoe Bay have been presented by Wheeler (S).<br />

The past occurrence of break-up by a given period was determined for each of the<br />

20 sites in the study. Break-up was defined as the first time the ice<br />

concentration dropped below S oktas (S/S ice cover) following the winter season.<br />

The results are presented in probalistic terms in Table 2. For example, there is<br />

a 10% probability that break-up will occur by at least the 4th period (June 19 to<br />

24) at Site 1. The occurrence of the earliest and latest observed break-up are<br />

also given.<br />

Similar statistics on the first occurrence of open-water and freeze-up are presented<br />

in Table 2. The first occurrence of open-water was defined as the first time the<br />

ice concentration dropped below 1 okta (l/S ice cover) following the winter season.<br />

Freeze-up was defined as the first-time the ice concentration equalled S oktas<br />

and remained S oktas until the twenty-fifth period or until data for that year were<br />

no longer available.<br />

The probability of having a given number of continuous and total open-water six-day<br />

periods was also <strong>com</strong>puted for each of the 20 sites. The number of continuous<br />

open-water periods was defined as the maximum number of consecutive open-water<br />

periods without any intervening pack ice invasions. The total number of open-water<br />

periods included all the open-water periods during the summer regardless of number<br />

of pack ice invasions. These results are summarized in Table 3. For example, at<br />

Site 1, there is a 50% probability of having 3 or more continuous open-water periods<br />

and 5 or more total open-water periods during the summer.<br />

6. Discussion<br />

Examining the 50% probability data in Table 2, it appears that on the average<br />

break-up in the study area takes place during the 6-th and 7-th periods, that is,<br />

between July 1 and 12. The ice first begins to break-up offshore and about a week<br />

later begins to break-up along the coast and inside the barrier islands. In general,<br />

break-up in the area has been observed to have occurred as early as the middle of<br />

June and as late as the beginning of August.<br />

The first appearance of open-water in the study area is more variable. The 50%<br />

probability data in Table 2 indicate that, on the average, open-water first appears<br />

close to the coast between July 7 and IS. Farther offshore the first occurrence<br />

of open-water does not usually take place until August 12 to 23, about a month<br />

later. Even though the ice first begins to break-up offshore, ice deterioration<br />

is accelerated near the coast due to river over-flooding of the sea ice. Both<br />

water and sediment on the ice surface reduces the ice albedo and enhances melting.<br />

There also appears to be a tendency to have open-water offshore (close to the<br />

20-meter isobath) first in the eastern part of the study area. The earliest<br />

open-water has been observed near the coastal sites is between June 25 and 30.<br />

It should be emphasized that even near the coast there are about one in ten years<br />

with no open-water and, farther offshore near the 20-meter isobath, as many as<br />

one in four years may have no open-water.<br />

The ice cover usually freezes over and remains in tact by the 22-nd and 23-rd<br />

periods. Freeze-up first takes place between October 5 and 10 near the coast and<br />

about a week later offshore. In general, the earliest freeze-ups have occurred<br />

between September 17 and ::lS, and about one in four years, freeze-up is not<br />

permanent until after October 2S.<br />

805


00 PROBABILITY OF OCCURRENCE<br />

0<br />

a-<br />

BREAK-UP FIRST OPEN-WATER FREEZE-UP<br />

SITE 10% 50% 90% E L 10% 50% 90% E L 10% 50% N-F E<br />

1 4 6 10 4 10 9 13 18 9 9%* 20 22 25% 20<br />

2 4 7 9 5 10 7 8 11 5 14 20 22 12% 20<br />

3 4 6 10 4 11 11 16 9 20%* 20 23 27% 20<br />

4 5 7 10 3 10 9 13 19 7 9%* 20 23 27% 20<br />

5 5 6 8 5 10 5 7 9 5 11 20 22 11% 20<br />

6 4 6 10 4 11 10 14 9 27% 20 23 25% 20<br />

7 5 7 9 5 10 6 8 10 5 13 20 22 17% 20<br />

8 5 6 9 3 10 5 7 9 5 11 20 22 12% 20<br />

9 5 7 9 4 10 9 12 18 8 9%* 20 23 22% 20<br />

10 5 7 9 4 10 8 11 8 12%* 21 23 22% 20<br />

11 1 6 9 1 10 11 15 5 15%* 20 22 27% 19<br />

12 5 7 9 5 10 7 9 11 6 13 20 22 12% 20<br />

13 6 7 9 5 10 8 10 15 5 9%* 20 22 20% 20<br />

14 4 7 9 4 10 8 11 15 5 5%* 20 22 20% 20<br />

15 2 6 9 1 10 9 15 5 19%* 20 22 27% 19<br />

16 5 7 9 5 10 5 8 11 5 15 21 22 12% 19<br />

17 4 7 9 3 10 9 12 5 15%* 20 23 22% 19<br />

18 3 6 9 2 9 9 13 20 5 9%* 19 23 22% 19<br />

19 2 6 9 2 10 8 14 20 5 9%* 19 23 22% 18<br />

20 5 7 9 4 9 7 10 14 7 4%* 20 22 12% 20<br />

10% E Earliest<br />

50% Percent Probability L Latest<br />

90% * Percent of years with no open-water<br />

N-F Percent of years with freeze-up after the 25th period<br />

Table 2: Period of break-up, first open-water, and freeze-up at different probability levels.


PROBAILITY OF OCCURRENCE<br />

CONTINUOUS OPEN-WATER TOTAL OPEN-WATER<br />

SITE 10% 50% 90% 10% 50% 90%<br />

1 10 3 1 11 5 1<br />

2 16 10 4 16 11 7<br />

3 10 4 0 10 4 0<br />

4 11 3 1 12 4 1<br />

5 15 12 5 15 12 8<br />

6 10 3 0 10 5 0<br />

7 15 7 3 15 11 7<br />

8 15 12 5 15 12 9<br />

9 12 5 1 12 6 1<br />

10 12 4 0 14 6 0<br />

11 11 3 0 11 4 0<br />

12 14 7 2 15 9 5<br />

13 13 4 1 13 7 1<br />

14 12 4 1 14 6 3<br />

15 9 3 0 10 4 0<br />

16 14 7 4 14 10 5<br />

17 10 3 0 11 5 0<br />

18 11 4 1 11 5 1<br />

19 11 4 0 12 5 0<br />

20 12 5 1 13 7 1<br />

Table 3: Number of 6-day periods with open-water<br />

at different probability levels.<br />

Data on the number of open-water periods in Table 3 show that, on the average,<br />

there are about SO to 70 days of open-water close to the coast and inside the<br />

barrier islands. As one moves offshore to more exposed areas inside the 20-meter<br />

isobth, the number of open-water days rapidly decreases to about 20 to 30 days.<br />

There are even fewer consecutive days of open-water without pack ice invasions.<br />

At the 90% probability level, we only have 30 to SO days of open-water close to<br />

the coast and up to 6 days offshore.<br />

The variability and limited number of open-water days offshore have a serious<br />

impact of offshore arctic petroleum operations in this area. Due to the variability<br />

in break-up and freeze-up dates, scheduling both winter and summer operations is<br />

difficult. It is necessary to begin an operation early while recognizing that<br />

the operation may not even get off the ground. Early mobilization of equipment<br />

and personnel and delays increase operating costs.<br />

The limited number of open-water days and time lost due to equipment failures and<br />

storms may require that projects, such as artificial gravel island construction, be<br />

conducted over several seasons. A fill island that can be constructed in 30 to SO<br />

days in MacKenzie Bay (9) may take two or more years to build in the Prudhoe Bay<br />

area. Winter standby costs will further increase the cost of construction.<br />

807


1.0 Introduction<br />

The successful design and operation of marine structures is highly<br />

dependent on a <strong>com</strong>plete knowledge of the wave climate. The term "marine<br />

structures" is used here to include coastal works and offshore platforms,<br />

as well as vessels, ranging from sailing boats to VLCC's. Since the<br />

design and operation of these structures relies heavily on the results<br />

of physical model experiments, hydraulics laboratories and ship towing<br />

tanks can be expected to be equally interested in having more <strong>com</strong>plete<br />

definitions of the wave climate available, as an input to their model<br />

studies.<br />

Unfortunately similar statements have been made by many other<br />

authors, at many times during the last few decades, but the number of<br />

structures which have failed, or the number of vessels which have capsized,<br />

due to the occurrence of "unexpected wave conditions", is still<br />

much larger than the total number of pleas which have been made to<br />

increase and to improve the measurement systems of wind generated water<br />

waves.<br />

Even today there is no area of the world's oceans, where adequate<br />

observations of the sea state are being made, or have been made over<br />

sufficiently long periods of time, to provide a satisfactory estimate<br />

of the wave conditions. Some wave recording stations have indeed been<br />

operational for several decades and their records can be used to predict<br />

with a good degree of accuracy the wave heights and periods, which are<br />

expected to occur in the areas of those stations with a given frequency<br />

of occurrence. The safe design and operation of marine structures,<br />

however, requires a much more <strong>com</strong>plete knowledge of the wave climate<br />

than the "wave height" and the associated "period". The present trend<br />

of designing and building structures in depths of water well beyond the<br />

breaker zone, has led to the requirement of defining the design wave<br />

conditions in much greater detail than was considered necessary only<br />

ten or twenty years ago.<br />

Some of the additional parameters which are known to be important<br />

in the design of marine structures include the wave steepness, the<br />

asymmetry of wave profile, the joint distribution of wave heights and<br />

periods, the wave direction, the crest lengths of waves, wave grouping<br />

(amplitude and period) and the spectral shape. While some of these<br />

additional parameters can still be obtained from re-analysinghistorical,<br />

instrumented wave records, others first require the development of more<br />

sophisticated instrumentation packages.<br />

810


There exists therefore today an urgent need not only to enlarge<br />

significantly the extent of wave measuring programmes, but also to<br />

develop new instrumentation packages and more <strong>com</strong>prehensive analysis<br />

programmes, so that in the future design wave conditions can be predicted<br />

more <strong>com</strong>pletely and more reliably.<br />

The following presentation will trace the development of some of<br />

the above listed factors and consider suggested definitions or techniques<br />

aimed at improving the analysis of wave records.<br />

2.0 Wave Modelling Techniques<br />

Much of the need for a more <strong>com</strong>plete description of the wave climate<br />

is a direct result of laboratory experiments, particularly of model<br />

tests of large, deep water structures. The use of irregular waves has<br />

identified many factors, other than the wave height and period, which<br />

affect the design and operation of marine structures.<br />

Variable speed, electric motors, traditionally used to drive wave<br />

boards in laboratory wave tanks or basins, have been replaced by much<br />

more versatile and powerful hydraulic actuators. The use of fast,<br />

digital <strong>com</strong>puters to control these hydraulic systems has made it possible<br />

to reproduce even the most <strong>com</strong>plex wave patterns. Indeed, the problem<br />

of producing realistic model sea states has been moved from the laboratory<br />

to nature. It is now largely the lack of a more <strong>com</strong>plete definition<br />

of the ocean wave conditions which limits the development of techniques<br />

to simulate accurately ocean sea states.<br />

A survey in 1979 of some 250 institutes [11 indicated that about<br />

50% of the 150 respondents to a questionnaire on wave generation and<br />

analysis techniques, reported irregular wave generating facilities.<br />

Interestingly, the hydraulics laboratories reported a considerably higher<br />

utilization of their irregular wave equipment than the ship towing tanks.<br />

This difference is probably partly due to the different nature of ship<br />

model testing, mostly a fairly linear process, as <strong>com</strong>pared to the testing<br />

of coastal and offshore structures, or for that matter the behaviour<br />

of large, moored vessels, where the non-linear effects be<strong>com</strong>e very important.<br />

It may, however, also be because of a basic difference in the<br />

philosophy of wave model testing, the deterministic versus the random<br />

approach. Johnson and Takezawa [21 review the various methods presently<br />

used, in a state-of-the-art paper at this years' International Towing<br />

Tank Conference.<br />

Whatever methods are used to generate irregular waves in a laboratory<br />

tank, the design of marine structures has benefitted greatly from<br />

811


To obtain a direct measurement of the wave steepness requires a more<br />

sophisticated sensor than is presently available. The ongoing development<br />

work to produce a sensor capable of determining wave direction will<br />

probably also be able to record the wave steepness directly.<br />

Kjeldsen further makes a plea for using the zero-down crossing<br />

analysis, rather than the zero-up crossing method, arguing that the<br />

zero-down crossing analysis produces a wave height which is physically<br />

more relevant, in particular with regard to the capsizing of vessels<br />

and shock pressures on structures.<br />

Another example of the impact of using irregular waves can be found<br />

in the effect of wave grouping on the design of structures. The phenomenon<br />

of wave grouping was known to affect moored vessels, harbours or<br />

other coastal structures with a natural response period close to the<br />

wave grouping period. More recently, laboratory wave tests have confirmed<br />

that the reliability of rubble mound breakwaters is also greatly<br />

affected by the occurrence of wave grouping [6,7]. Again, by means of<br />

laboratory experiments, new parameters have been defined, capable of<br />

describing the sea state more <strong>com</strong>pletely.<br />

The Danish Hydraulics Institute has reported the results of experiments<br />

using short-crested waves [8], and their effects on the mooring<br />

forces of vessels. They concluded that in order to evaluate the feasibility<br />

of using exposed offshore mooring platforms, it is essential to<br />

use a three-dimensional wave generation system, which simulates the<br />

short-crested waves. No wave measuring system is presently available<br />

to provide prototype information of the crest lengths of wind generated<br />

water waves.<br />

Much improvement is already possible, however, by more extensive<br />

analysis procedures of instrumented wave records. The routine reporting<br />

of additional parameters will contribute significantly to a better<br />

understanding of the required input wave conditions for model tests. As<br />

soon as possible, international agreement should be reached on the definition<br />

of such additional parameters and which of these to include in<br />

the normal wave climate reporting procedures. Only the most important<br />

of these have been included in the following chapters.<br />

3.0 Wave Heights and Wave Periods<br />

The selection of input wave conditions for the design of marine<br />

structures requires a knowledge of the short term statistics, as well<br />

as the long term statistics or extreme values of the various parameters<br />

defining the wave climate. Many papers and books have been written<br />

813


on both subjects, but today there still exists a great deal of contro­<br />

versy on which method to use.<br />

The long term statistic methods fall into two categories. The<br />

simplest procedure, mostly used by engineers, is to plot all available<br />

measured data on some carefully selected graph paper, so that the data<br />

will most closely follow a straight line. Commonly used probability<br />

laws on which the scales of graph paper are based, are log-normal,<br />

Weibull, Gumbel or Rayleigh distributions. Lately, the Weibull distri­<br />

bution appears to find the greatest following and several studies of<br />

long periods of wave records have indicated good fits.<br />

The second method of determining extreme values involves the use<br />

of a model, which has been calibrated for a data base of recorded wave<br />

parameters. The most frequently used approach is the hindcasting technique;<br />

some of the recent models in this method involve reasonablyaccurate<br />

descriptions of the physical wave generation process. Both methods<br />

have their problems and limitations. Common to both, is the problem<br />

that most data bases are too short and that often the available data<br />

are from a different population than for which the extreme values are<br />

required.<br />

The short term features of wave conditions involve not only the<br />

distribution of individual wave heights within a storm, but also of<br />

course the distribution of wave periods, jointly with the wave heights,<br />

the wave steepness as already mentioned earlier, wave grouping phenomena<br />

and spectral shapes.<br />

Longuet-Higgins paper in 1952 [9] on the distribution of the<br />

heights of sea waves has be<strong>com</strong>e a classic. He derived that the probability<br />

density of the wave height is given by the Rayleigh distribution.<br />

He also discussed in various subsequent papers the distribution of wave<br />

periods and in 1975 published a paper on the joint distribution of wave<br />

periods and wave heights [10]. Fig. 2 shows graphically this joint distribution.<br />

Longuet-Higgins and several others have <strong>com</strong>pared the theo­<br />

retical curves with recorded data sets for various parts of the ocean<br />

and in general good agreements have been found.<br />

4.0 Spectral Shapes<br />

The use of irregular waves for model testing is largely based on<br />

a spectral input, although this method has some serious limitations.<br />

The two most widely used spectral formulations are the Pierson-Moskowitz<br />

[11] and the Jonswap spectra [12].<br />

8"


or in words: GF is the standard deviation of the SIWEH about its mean<br />

and normalized with respect to this mean. Fig. 3 illustrates some<br />

examples of wave trains with a <strong>com</strong>mon spectral density function, but<br />

different grouping factors 1 also shown are the SIWEH spectra for these<br />

wave trains, which indicate substantial differences.<br />

These proposed formulations for describing wave grouping are find­<br />

ing a certain amount of support from other institutes [16].<br />

The definition of the grouping factor, using the SIWEH-spectral<br />

density function does not only provide a valuable new parameter to des­<br />

cribe the wave conditions, but it can also be used to generate sea<br />

states in laboratory wave tanks which have the same degree of wave group­<br />

ing as measured in a particular area. This is most important to improve<br />

the design and operation of marine structures.<br />

6.0 Conclusions<br />

The increasing requirement for designing and building structures<br />

in deep water has led to the need for a more <strong>com</strong>plete definition of the<br />

wave climate. Much work has been done over the past few years to<br />

develop new analysis techniques and to formulate new models or parame­<br />

ters. The most urgent requirement at the present time appears to be<br />

reaching international agreement on a new set of parameters, which<br />

should be included routinely in all wave data analysis programmes. The<br />

sooner such an agreement is reached, the earlier new data bases can be<br />

developed, either from existing wave records or from new recordings,<br />

which can then be used to develop short and long term statistics of<br />

these new parameters.<br />

References<br />

1. Ploeg, J. and Funke, E.R., "A Survey of 'Random' Wave Generation<br />

Techniques", Proc. Seventeenth Coastal Engineering Conference,<br />

Sydney, Australia, 1980.<br />

2. Johnson, B. and Takezawa,S., "State of the Art Review of Irregular<br />

Wave Generation and Analysis", 16th International Towing Tank<br />

Conference, 1981.<br />

3. Funke, E.R. and Mansard, E.P.D., "SPLSH - A Program for the Synthe­<br />

sis of Episodic Waves", National Research Council, Hydraulics<br />

Laboratory Technical Report LTR-HY-65, 1979.<br />

4. Kjeldsen, S.P. and Myrhaug, D., "Interaction and Breaking of<br />

818<br />

Gravity Water Waves in Deep Water", River and Harbour Laboratory,<br />

Trondheim, Norway, 1978.


5. Kjeldsen, S.P. and Myrhaug,D., "Kinematics and Dynamics of Breaking<br />

Waves", River and Harbour Laboratory, Trondheim, Norway, 1975.<br />

6. Johnson, R.R. et aI, "Effects of Wave Grouping on Breakwater<br />

Stability", 16th Coastal Engineering Conference, Hamburg, Germany,<br />

1975.<br />

7. Goda, Y., "Numerical Experiments on Wave Statistics with Spectral<br />

Simulation", Port and Harbour Research Institute, 1970.<br />

S. Kirkegaard, J. et aI, "Effects of Directional Sea in Model Testing",<br />

ASCE Specialty Conference, Ports 'SO, 19S0.<br />

9. Longuet-Higgins, M.S., "On the Statistical Distribution of the<br />

Heights of Sea Waves", Journal of Marine Research, 1952.<br />

10. Longuet-Higgins, M.S., "On the Joint Distribution of the Periods<br />

and Amplitudes of Sea Waves", Journal of Geophysical Research,<br />

1975.<br />

11. Pierson, W.J. and Moskowitz, L., "A Proposed Spectral Form for<br />

Fully Developed Wind Seas Based on the Similarity Theory of<br />

Kitaigorodskii", Journal of Geophysical Research, 1964.<br />

12. Hasselman, K. et aI, "Measurements of Wind-Wave Growth and Swell<br />

Decay during JONSWAP", Deutsche Hydrographische Zeitschift, 1973.<br />

13. Thompson, E.F., "Energy Spectra in Shallow U.S. Coastal Waters",<br />

Coastal Engineering Research Centre, Fort Belvoir, 19S0.<br />

14. Funke, E.R. and Mansard, E.P.D., "Synthesis of Realistic Sea States<br />

in a Laboratory Flume", National Research Council, Hydraulics<br />

Laboratory Technical Report LTR-HY-66, 1979.<br />

15. Funke, E.R. and Mansard, E.P.D., "'Random' Wave Generation by<br />

Means of a Generalized Computer Software System", Proc. XIX IAHR<br />

Congress, New Delhi, India, 19S1.<br />

16. Houmb, O.G., "On the Wave Climate of the North Sea and the Problems<br />

of Determining Design and Operational Conditions for this Area",<br />

Wave Information Workshop, Bedford Institute of Oceanography,<br />

Halifax, 19S0.<br />

S19


Mass transportation of water in the direction of wave propagation is<br />

an inherent feature of wave action and Wiegel and Johnson provide the<br />

following expression for its velocity:<br />

V max<br />

in which z is the water depth (rated negative from the water level<br />

downwards). This expression has been substantiated by observations [2J.<br />

,-<br />

Since "H/L equals the maximum gradient of the water surface it<br />

follows that the velocity of mass transportation will decrease rapidly<br />

with decreasing wave steepness.<br />

The theoretical upper limit of the H/L ratio equals 0,14 or 1/7 as<br />

determined by Mitchel and Havelock and quoted by Wiegel and Johnson [2J.<br />

When testing with plunger type wave machines it was not possible to<br />

produce waves of such steepness but waves with H/L = 1/10 could be<br />

produced.<br />

In the wave period range of 1.5 to 2.5 seconds waves of steepness 1/10<br />

produce surface mass transport of more than 1,000 meters per hour and<br />

waves of steepness 1/20 produce 200 to 300 meters per hour. In<br />

<strong>com</strong>parison swell waves of this period when occurring in nature would<br />

have waves of steepness about 1/30 and provide surface mass transpor­<br />

tation of only 50 meters per hour which would be difficult to<br />

destinguish from random currents often present.<br />

Decay of wave energy and wave height due to fluid friction is very<br />

small and in the wave range being considered here would be insignificant.<br />

Wiegel and Johnson [2J.<br />

The viscosity of water increasingly charged with ice slush would<br />

increase but the resulting damping is still small for waves of any<br />

appreciable length.<br />

The wave pattern generated by a machine or several machines in line<br />

form a wave train. Wave height reduction is caused by diffraction of<br />

waves or spreading of the wave train during progression as it occurs<br />

when waves move through a gap into a sheltered region. Relative<br />

diffraction losses will be reduced when the breadth of wave generation<br />

is increased. We have found that a breadth of wave generation equal to<br />

approximately 6.5 L will suffice to provide a far-reaching wave train<br />

as also confirmed by calculation procedures.<br />

822<br />

(2)


with wave machine breadth equal to 6.5 L, the wave height reduction<br />

along centre line will reach 0.6 H at a distance from the generator of<br />

approximately 75 L but further reduction of wave height down to 0.5 H<br />

will take place only very slowly over the next 1000 L so that one half<br />

wave height will be mainlained over a very long distance.<br />

3. Suspended Ice in the Wave Train<br />

In agitated water the initial ice formations, discoids, are very small<br />

but from these will grow the needle - like fragments or spicules<br />

recognized as frazil ice. Masses of spicules form slush ice.<br />

Due to the relatively high deformation drag associated with the motion<br />

of small particles in a viscous fluid, the rising velocity of the<br />

smallest ice pieces are negligible in <strong>com</strong>parison with the mixing<br />

currents and it is only the larger ice pieces that tend to accumulate<br />

at surface due to buoyancy forces. For this reason ice may be distri­<br />

buted throughout a surface layer of considerable thickness.<br />

Studies of ice formation in water that has been pre-cooled to the<br />

freezing point temperature and then subjected to surface cooling and<br />

wave action was carried out in conrection with earlier wave tests as<br />

discussed in "Ice Free Harbours" [1] in which primitive tests of rough<br />

accuracy are discussed. The amount of ice formed under wave action was<br />

found to be of the order of 250 grammes/hour per square meter of<br />

surface for each 1 0 Celcius that the air temperature is below the<br />

freezing point.<br />

4. Ice Transportation Capacity of the Wave Train<br />

The capacity of the wave train for transporting surface formed ice<br />

would depend not only on the velocity of its transportation but also<br />

on the capacity of the surface waters for holding formed ice in<br />

suspension. As the amount of ice in the wave train grows, a greater<br />

proportion of it would congregate near the surface and a layer<br />

extending from surface to the level where the energy level has been<br />

halved coinciding with the halving of the mass transportation velocity<br />

would contain a great majority of the suspended ice.<br />

In order to obtain a practical basis for <strong>com</strong>paring wave trains of<br />

different make-up we have chosen to define this layer thickness as the<br />

"active layer" of the wave train and to define as the "Reach" of<br />

823


the wave train the distance covered by the mass transport by the time<br />

the ice content of the active layer has reached ten percent (10% ice<br />

spicules and go% water).<br />

The active layer thickness would by this definition be derived from<br />

and be<strong>com</strong>es<br />

e4fiz/L = 0.5<br />

Z = -0.055L<br />

a<br />

The surface velocity of the mass transportation according to (2) will<br />

be decreasing along the wave train towards<br />

_ 1 2<br />

V s = ( II 20) x C = 0.025 C<br />

and average velocity of the mass transportation within the limits of<br />

the active layer can be estimated as<br />

Va = 0.02 C<br />

Let RK denote the "Reach" in meters of the wave train at a given<br />

design temperature K in 0 Celcius and let t denote the time variant<br />

in hours, then<br />

(3)<br />

RK V<br />

a<br />

x t x 3600 72 x t<br />

and -K 250 t 0.1 10 6<br />

x x = x Z<br />

a<br />

x<br />

or<br />

RK<br />

1580<br />

L C<br />

=K<br />

(meters) (4)<br />

Since basic concepts regarding ice contents and layer thickness were<br />

arbitrarily chosen the expression (4) should only be used for the<br />

purpose of <strong>com</strong>paring wave trains. Absolute expressions of capabilities<br />

of waves can only be developed with increased knowledge of ice<br />

behavior in ,waves.<br />

5. Practicable Wave Machines<br />

Wave machine installations for ice prevention in harbours, channels,<br />

navigation lock approaches, stagnant waters etc. must be designed to<br />

suit basic requirements including<br />

Power efficiency and automation<br />

Infrequent and easy maintenance<br />

Survival against environmental forces<br />

Model tests with floating wave machines were carried out several years<br />

ago as discussed in "Ice Free Harbours" llJ. Based on one of the models<br />

824


6. Ice Removal by Wave Train, Tabulation<br />

The table of wave trains is based on calculated results and generally<br />

accepted formula and data. The column REACH is prepared to <strong>com</strong>pare the<br />

ice clearing capacities of different wave trains under Arctic condi­<br />

tions but should not be considered as an absolute measure of these<br />

capacities.<br />

x xx xxx<br />

Wave Characteristics Wave Train Wave Machine Installation<br />

Period Velo Length Height Mean REACH Com- Dis- Horse<br />

-city Trans ato bined place -power<br />

-port -30 C Breadth -ment<br />

x<br />

Sec mlsec m m mlhour m m Tonnes HP<br />

1.5 2.34 3,51 0.35 168 433 3x 8m 10 T<br />

1. 75 2.73 4.78 0.48 196 689 3x12m 28 T<br />

2.0 3.12 6.24 0.62 225 1,030 3x12m 48 T<br />

2.25 3.28 7.37 0.74 236 1,280 4x12m 95 T<br />

2.5 3.90 9.75 0.98 281 2,010 3x20m 193 T<br />

3.0---- 4.68 14.0 1.40 337 3,470 3xl30m 600 T<br />

9 HP<br />

25 HP<br />

45 HP<br />

90 HP<br />

260 HP<br />

980 HP<br />

3.5 5.46 19.1 1.91 393 5,510 3x40m 1480 T 2700 HP<br />

4.0 6.24 25.0 2.50 449 9,870 3x50m 3160 T 6400 HP<br />

The table is based on all waves starting out with the same original<br />

steepness of H/L = 1/10<br />

xx Mass transportation figures represent the calculated mass transport<br />

velocity in the wave train at a distance from the wave generators of<br />

at least 100 L and a depth below surface of 0.055 L.<br />

The REACH is based on _2 0 C cold sea water and a continuing air tempera­<br />

tur of -30 0 C and represents the distance from the wave machine covered<br />

by the mass transport before or by the time the ice content of the sur­<br />

face waters has reached ten percent (10% ice spicules and 90% water).<br />

xxx A wave machine installation as planned may consist of several,<br />

usually three, individual machines each with its own engine, but,<br />

alligned and linked together and attached to the sea bottom. The Dis­<br />

placement Tonnages represent the weight of equipment, pontoons and<br />

machinery of all the wave machine installation. The horsepower figures<br />

represent the <strong>com</strong>bined power rating of all of the engines of the<br />

installa ticn.<br />

827


eventually have to cope with either<br />

Fast ice<br />

Open pack ice<br />

Closed pack ice<br />

The first two conditions hold no problems because the wavetrain can<br />

maintain an open channel through stationary ice and the disposal rates<br />

for the slush ice amongst the ice floes at sea are relatively small.<br />

However, icebreaker service will be needed to maintain access if<br />

closed pack ice moves across the harbour antrance.<br />

It is estimated that inflow of ice into the harbour due to moderate<br />

tidal inflow can be held at bay by the disposal wave train because<br />

Doppler effect will steepen the waves against counter flow.<br />

8. References<br />

1 Andersen, Per F. (1972) "Ice Free Harbours" The Engineering<br />

Journal, July-August 1972. The Journal of the Engineering<br />

Institute of Canada.<br />

2 Wiegel, R.L. and Johnson, J.W. (1951) "Elements of Wave Theory"<br />

<strong>Proceedings</strong> First Conference Coastal Engineering, Long Beach,<br />

California.<br />

829


WIllDS AIID WAVES III LAllCASTD SOUIID<br />

A11IOSPIIBRIC DVIIlOIIIIBlIIT SnVICE DOiflfSVIBW, OIlTARIO,<br />

CAIIADA<br />

This paper presents an overview of a derived wind and wave climatology of<br />

Lancaster Sound in the Canadian Arctic.<br />

Hourly wind speeds were synthesized at three locations in Lancaster Sound.<br />

The geostrophic equations were used to obtain wind data from 33 years of gridded<br />

6-hourly sea-level pressure data <strong>com</strong>piled by the Fleet Numerical Oceanography<br />

Center in Monterey, California. Hourly wind values were obtained by linear inter­<br />

polation of the 6-hourly 'u' and 'v ' <strong>com</strong>ponents of the pressure gradient. The 21-<br />

year wave hindcasts were based on the Bretschneider equations using geographical<br />

features and average minimum monthly ice conditions to limit the fetch.<br />

Statistics of the wind and wave climatologies are presented in terms of fre­<br />

quency distributions and extremes.<br />

There are obvious drawbacks in using gridded pressure as a data base. Many of<br />

these drawbacks are identified and some are examined in detail. As part of this<br />

examination, <strong>com</strong>parisons of derived data versus real data at Resolute and Ocean<br />

Weather Station Bravo are made and case studies of particular significant storms<br />

in the Lancaster Sound - Baffin Bay area are discussed.<br />

830


1. BACGIlOIIIID<br />

The increasing interest in offshore oil and gas exploration during the 1970's<br />

and 1980' s has led to a nlDDber of studies which seek to determine the possible<br />

hazards to and from the environment associated with exploration activities. One of<br />

the areas currently of high interest is Lancaster Sound.<br />

Lancaster Sound is located between Devon Island and northern Baffin Island<br />

and connects with Baffin Bay at its eastern end (see Figure 1). It is one of a<br />

series of sounds and straits known collectively as Parry Channel, this latter be­<br />

ing a proposed shipping route for Liquid Natural Gas (LNG) [I].<br />

A study performed in 1977 [2] looked in a preliminary fashion, at different<br />

meteorological and oceanographic facets of Lancaster Sound. Similar techniques<br />

were used in a parallel study of northwestern Baffin Bay [3]. The present study<br />

concentrates on winds and waves in Lancaster Sound.<br />

2. DATA SOUItCBS<br />

The wind and wave data used in this study were derived from a set of gridded<br />

6-hourly hemispherical pressure maps <strong>com</strong>piled by the Fleet Numerical Oceanography<br />

Center in Monterey, California. The grid interval is 381 km (see Figure 1). The<br />

gridded pressure data began in January 1946 and ended in December 1978.<br />

Three locations in Lancaster Sound were chosen for calculating winds and<br />

waves. They are 77°W, 83°W and 89°W all three along the latitude 74.25°N, and will<br />

henceforth be referred to as positions A, Band C respectively (see Figure 1). As<br />

well, winds were <strong>com</strong>puted at 56.5°N, 51.0oW and at 74.7°N, 95.0 o W which correspond<br />

to Ocean Weather Station "Bravo" and to Resolute respectively. In the tables,<br />

positions SB and R will refer to <strong>com</strong>puted data while OWSB and YRB will refer to<br />

observed data for "Bravo" and Resolute respectively.<br />

The Summary of Synoptic Meteorological Observations (SSMO) tables <strong>com</strong>piled to<br />

1972 for the Lancaster Sound - Northwestern Baffin Bay and OWSB were used for ver­<br />

ification of the <strong>com</strong>puted winds and waves for the marine areas. The Atmospheric<br />

Environment Service data archives were used to <strong>com</strong>pare winds <strong>com</strong>puted for<br />

Resolute.<br />

831


Individual cases were assessed using information <strong>com</strong>piled by Thomson and<br />

Vickers ([4), and personal <strong>com</strong>munication) on the 81 gale force storms in North­<br />

western Baffin Bay over the period 1969-79.<br />

Winds at a specific location were <strong>com</strong>puted using Bessel's function in cubic<br />

form. A Sixteen-point grid array was used to obtain 'u' and 'v' gradients at a<br />

specific location.<br />

Wind speed and direction were obtained from the u and v <strong>com</strong>ponents. Hourly<br />

values were obtained by linear interpolation of the u and v <strong>com</strong>ponents. No inter­<br />

polation was done if four or more 6-hourly consecutive maps were missing. This<br />

last condition only occurred three times in 33 years, none of which were in the<br />

June to October period.<br />

4. WAVE COIIPOTATIOIIS<br />

Hourly wave height and period data were produced under contract by Hydrotech­<br />

no logy Limited. Due to initial difficulties with the gridded pressure data and to<br />

time constraints, the wave data were <strong>com</strong>puted for the years 1956-71 and 1974-78,<br />

for the five-month periods June to October. The wave hindcasts were based on the<br />

Bretschneider equations. Wind fetches were limited by geography and minimum ob­<br />

served ice conditions for each month (see Figure 2). No limitation was imposed due<br />

to isobaric curvature. Fetches were adjusted when coastal or channel topography<br />

warranted changes.<br />

5. VERIFICAnOli or PIlOCBDUUS<br />

a) Stoma<br />

Two specific storms are considered. One provides good verification of the<br />

gridded pressure data, the other provides poor verification. These storms are de­<br />

picted in Figures 3 and 4 where the lines of pressure are taken from a hand analy­<br />

sis of data and the grid values are indicated below the appropriate square. All<br />

values are in kiloPascals.<br />

832


• Pressure grid positions<br />

e Lancaster Sound sites<br />

Figure 1. Geographical reference map<br />

June - - -- August - ........... .<br />

September - all open water October<br />

Figure 2. Boundary of minimum ice edge by month<br />

for June to September


The November 24, 1976 storm (Figure 3) shows the same direction of the gradi­<br />

ent from both methods, but the hand analysis indicates a much stronger gradient in<br />

the north than does the numerical data. This particular double-low pressure case<br />

is unusual in that the northern low pressure was developed in Baffin Bay rather<br />

than advected from upstream.<br />

The September 26, 1974 storm (Figure 4) was also a notable one as described<br />

by Thomson and Vickers [4). "One of the most intense storms in the period 1969-79<br />

to affect the study area, this is a type of storm that advects warm air into Baf­<br />

fin Bay and produces strong southeasterly winds. The storm is all the more signi­<br />

ficant since at the time Baffin Bay was ice-free which allowed a long fetch length<br />

(840 km) for waves to build." In this case the numerical data agrees very closely<br />

with the hand analysis.<br />

b) Ceostrophic ViDds st Location A.<br />

A <strong>com</strong>parison of hand abstracted geostrophic winds (Thomson and Vickers, per­<br />

sonal <strong>com</strong>munication) and <strong>com</strong>puted values was carried out for 381 cases.<br />

Of the 381 pairs of wind direction at location A, 80% were within 20· of each<br />

other and 91% were within 30·. A difference of 30· between the two methods is<br />

viewed as being reasonably acceptable. Only 3% of the pairs had a deviation great­<br />

er than 50·.<br />

The wind speed did not fare as well. The hand abstracted winds varied from<br />

one-half to five times the values of the numerically <strong>com</strong>puted winds with an aver­<br />

age close to one and one-half times the numerical value. This result will be dis­<br />

cussed later.<br />

c) .. solute ViDds.<br />

Tables 1 and 2 give a <strong>com</strong>parison of winds recorded at Resolute (YRB) and<br />

winds <strong>com</strong>puted from the grid pressure data at position R (Resolute). Table 1 lists<br />

the percentage frequency of directions by month. The calm class does not appear in<br />

the <strong>com</strong>puted winds since the occurrence of a zero gradient is very rare in gridded<br />

data (4 calms in 37,956 cases). Table 2 lists the percentage frequency by 5 mls<br />

speed classes by month (o


Figure 3. Surface pressure analysis and numerical grid<br />

values for 1800 GMT,November 24, 1976.<br />

Figure 4. Surface pressure analysis and numerical grid<br />

values for 0000 GMT, September 26, 1974.


836<br />

TABLE 1<br />

PERCENTAGE FREQUENCY OF WINO DIRECTIONS<br />

AT RESOWn<br />

1953-1978<br />

JURE JULY AUG. SEPT. OCT. NOV.<br />

N YRa 22.6 17.2 18.0 31.5 25.6 21.8<br />

It 23.0 19.7 20.1 26.1 21.5 23.4<br />

NE YRB 8.2 7.0 5.2 10.3 10.0 9.6<br />

It 9.0 11.9 9.4 9.1 9.1 10.0<br />

E YRB 13.8 13.5 19.2 12.6 15.5 18.5<br />

It 8.6 8.8 12.6 7.7 7.8 9.0<br />

SE YRB 10.6 14.6 14.7 7.3 10.1 9.9<br />

It 4.6 6.2 11.3 6.1 9.2 7.5<br />

S YRB 8.0 9.1 9.0 7.0 7.9 6.3<br />

It 10.1 11.4 12.2 8.4 10.9 9.2<br />

SW YRB 2.2 1.5 1.8 3.0 3.1 1.7<br />

It 7.8 10.1 7.9 7.1 6.4 6.5<br />

W YRB 15.9 19.5 14.6 8.8 8.4 6.4<br />

It 13.9 15.0 1l.0 1l.8 lJ.O 11.1<br />

NW YRB 14.2 12.0 11.6 14.8 13.9 15.2<br />

It 23.0 16.9 15.7 23.9 22.1 23.3<br />

CAu{ YRB 4.6 5.7 6.0 4.7 5.6 10.6<br />

'It 0 0 0 0 0 0<br />

TABLE 2<br />

PERCENTAGE FREQUENCY OF WIRD SPEEDS<br />

AT RESOwn<br />

1953-1978<br />

_/8 JURE JULY AUG. SEPT. OCT. lIOI'.<br />

0-5 Yl!.8 45.5 50.1 47.3 37.8 41.3 53.2<br />

It 36.7 44.5 45.7 33.9 28.7 29.2<br />

5-10 YRB 39.0 36.6 37.0 43.3 37.5 29.8<br />

It 47.9 46.3 42.5 45.6 44.5 46.0<br />

10-15 YRB lJ.6 11.0 11.5 16.6 16.5 12.4<br />

It lJ.7 8.8 10.6 17.1 22.2 20.0<br />

15-20 YRB 1.8 2.1 3.5 2.1 4.4 3.9<br />

It 1.4 0.4 1.2 3.0 4.3 4.2<br />

20-25 YRB 0.1 0.2 0.4 0.2 0.3 0.5<br />

R 0.1 0 0.1 0.3 0.4 0.6<br />

25-30 YRB 0 0 0 0.03 0.06 O.lJ<br />

R 0.10 0 0 0.06 0.03 0.6<br />

30-35 Yl!.8 0 0 0 0 0 0.3<br />

It 0 0 0 0 0 0<br />

35-40 Yl!.8 0 0 0 0 0 0<br />

R 0 0 0 0 0 0


For both real and <strong>com</strong>puted data, 100% represents 3100 to 3200 values per<br />

month. Thus the 0.03% for wind speeds above 30 mls in Table 2 represents one oc­<br />

currence of such winds for each category, both of which occurred on November 20,<br />

1965. On this date the maximum observed wind was easterly at 35.8 mls and the<br />

maximum <strong>com</strong>puted wind was easterly at 24.7 m/s.<br />

Easterly to southeasterly winds at Resolute are known to be poor indicators<br />

of the strength of the pressure gradient due to local topography. In these condi­<br />

tions, observed winds tend to be greater than might normally be expected if ter­<br />

rain were not considered.<br />

d) Ocean weather Statio. Bravo.<br />

OWS Bravo was felt to hold the better opportunity for <strong>com</strong>paring the proce­<br />

dures used. Comparisons were primarily ac<strong>com</strong>plished by using "Percentage Exceed­<br />

ance" graphs.<br />

Figures 5 and 6 show the percentage exceedance of winds both observed at<br />

Bravo and <strong>com</strong>puted at Bravo from the pressure grid for August and October. Inset<br />

in each figure are tables of frequency of direction. The calm class for the SSKO<br />

statistics has not been considered. The wind speed plots are quite close and the<br />

wind direction percentages are more accurate than using persistence and a random<br />

distribution of winds which would allot 12.5% to each <strong>com</strong>pass direction.<br />

6. VALIDITY OF PROCBDUUS<br />

The different <strong>com</strong>parisons of the previous section demonstrate that the method<br />

used in this study does not yield absolute answers.<br />

Individual storms can be mis-analyzed by numerical techniques for <strong>com</strong>puting<br />

pressure grid maps. However one of the important aspects for wave generation in<br />

this topographically rugged area is wind direction. This parameter has been cap­<br />

tured fairly well by the numerical method since 91% of the wind cases <strong>com</strong>pared<br />

were within 30° of each other. Of the 3% of the cases where direction differed by<br />

more than 50°, half involved difficult pressure patterns in which a displacement<br />

of 50 to 100 km on the map would result in a totally different wind which would<br />

agree with the numerical wind within 30°; such a level of uncertainty can be con-<br />

837


30°; such a level of uncertainty can be considered acceptable in view of the fact<br />

that the grid spacing of the numerical data is 381 km.<br />

The large differences observed in the wind speeds for the case <strong>com</strong>parisons is<br />

understandable to a point in that numerically <strong>com</strong>puted winds were expected to be<br />

lower than a true geostrophic wind as a result of numerical smoothing techniques<br />

normally incorporated. Only 45 of the 381 cases yielded a numerical wind greater<br />

than the hand abstracted wind with the largest difference not exceeding 6 m/s.<br />

The winds obtained by numerical means for Resolute provided a good indication<br />

of conditions on a frequency basis for this location. The lack of terrain influen­<br />

ces in the numerical winds is evident in the E, SE, and SW direction which are the<br />

directions of the major topographical obstructions at Resolute.<br />

Where topography was not a factor, at ship Bravo, the results are encourag­<br />

ing. Computed winds were a good approximation of observed winds.<br />

7. LAllCASTBIl somm<br />

There are very few actual observations taken from ships in Lancaster Sound.<br />

The SSMO Tables summarize the available information to the end of 1972 on a grid<br />

square basis. Two such areas are pertinent to this study, areas 11, and 12 (see<br />

Figure 1). Area 11 has a total of 263 observations for the month of August and 230<br />

for September while area 12 has 158 for August, 143 for September and 66 for<br />

October.<br />

Figure 7 <strong>com</strong>pares the winds <strong>com</strong>puted for August at A with area 12 winds.<br />

Figure 8 <strong>com</strong>pares the waves <strong>com</strong>puted for August at A with area 12 waves. Frequency<br />

tables of the waves <strong>com</strong>puted at A and B are given in tables 3 and 4 respectively,<br />

as well as theoretical wave heights for specific return periods based on Gumbel<br />

statistics and their extremes (1.8 times the significant wave heights).<br />

8. DlSCUSSIOil<br />

The data presented for the three locations provide a first guess estimate of<br />

sea state conditions in Lancaster Sound. Due to the many variables involved in the<br />

procedure, too many qualifiers could be applied to these results at this time.<br />

839


One of the most significant variables pertinent to this area is ice cover.<br />

Minimum ice conditions were specified in this study in order to obtain results<br />

which would approximate the higher extremes of the wave climatology of the area.<br />

The results obtained appear to provide a reasonable estimate, slightly on the high<br />

side, of actual conditions encountered in the area.<br />

Validation of techniques by using ship information is a questionable method<br />

given the sparseness of information. Only long-term, continuous information from a<br />

fixed location can remedy this situation.<br />

REFEBElfCES<br />

1. Smiley, B.D. and A.R. Milne, 1979: LNG Transport in Parry Channel: Possible<br />

environmental hazards. Institute of Ocean Sciences, Patricia Bay,<br />

Sydney, B.C. 47 pp.<br />

2. Duck, P.J., M.O. Berry and D.W. Phillips, 1977: A Preliminary Analysis of<br />

Weather and Weather-Related Factors in Lancaster Sound. Unpublished<br />

Manuscript. Project Report No. 32, Meteorological Applications Branch,<br />

Atmospheric Environment Service, Downsview, Ont. 29 pp.<br />

3. Maxwell, J.B., P.J. Duck, R.B. Thomson and G.G. Vickers, 1980: The Climate of<br />

Northwestern Baffin Bay. Unpublished Manuscript. Canadian Climate Centre<br />

Report No. 80-2. Atmospheric Environment Service, Downsview, Ont. 10Spp.<br />

4. Thomson, R.B. and G.G. Vickers, 1980: Extreme Storm Study of Northwestern<br />

842<br />

Baffin Bay. To be published by Atmospheric Environment Service, Downs­<br />

view, Ont. 28 pp.


D. Carter, D.Sc.<br />

Y. Ouellet, D.Sc.<br />

P. Pay, P. Eng.<br />

Abstract<br />

FRACTURE OF A SOLID ICE COVER BY<br />

WIND-INDUCED OR SHIP-GENERATED WAVES.<br />

Consultant, 1281 Bishop, Quebec.<br />

Professor, Universite Laval.<br />

Transport Canada, Waterways Development<br />

Canada<br />

Canada<br />

Ottawa<br />

The paper presents the basic equations governing the propagation of waves in<br />

ice-covered waters of any arbitrary finite depth. Practical relationships are also<br />

proposed to easily evaluate the modifications of the wave characteristics while<br />

entering the ice-covered waters, and the minimum wave amplitude necessary for the<br />

fracture of the ice bordering the open waters.<br />

Resume<br />

La presente <strong>com</strong>munication rappelle les equations de base qui gouvernent la<br />

propagation des vagues dans des eaux, de profondeur arbitraire, recouvertes de<br />

glace. Des relations pratiques sont aussi proposees pour permettre une evaluation<br />

facile des modifications subies par les vagues lors de leur entree dans les eaux<br />

recouvertes de glace ainsi que l'amplitude suffisante pour briser la glace qui<br />

borde les eaux libres.<br />

Introduction<br />

In the St. Lawrence ship channel, large pieces of border ice are often observed,<br />

during winter, breaking off through the action of wind-induced or ship-generated<br />

waves. The broken ice floes, not only, cause hazardous conditions for shipping but<br />

frequently telescop and <strong>com</strong>press to form heavy ice jams and packs, with subsequent<br />

sharp increase in upstream water levels.<br />

The present study will aim, first, at determining the modifications undergone by<br />

the waves that arrive at the ice edge and propagate into the ice-covered waters.<br />

Then, we will try to estimate the minimum wave amplitude necessary to cause fracture<br />

of bordering ice cover.<br />

843


where the sinusoidal ice displacement is given by:<br />

n<br />

After lengthy algebraical manipulations, we finally obtain:<br />

In open water, the celerity, Co' of a surface wave, including surface tension<br />

effects, may be expressed by the following equation (Lamb 1932):<br />

k s<br />

C =_0_<br />

o [ P<br />

+ -L tanh (k d)]<br />

k 0<br />

o<br />

!<br />

Figure 1 summarizes how C i varies with wave period for different ice thicknesses<br />

and water depths. It can be seen that for long period wave the phase velocities in<br />

ice-covered and open waters are practically the same. All graphs in this paper were<br />

plotted with E = 6 x 10 9 N I m 2 , a = 15 x 10 5 N I m 2 and \I - 0,3, representative<br />

values for natural ice covers, (Carter, 1970).<br />

1.2 - Wave Length<br />

By considering that the wave impinging on the ice edge acts as a forcing term in<br />

the equation of motion of the ice, then we may argue that the edge of the ice responds<br />

essentially at a frequency corresponding to the period of the incident wave. Thus,<br />

by this simple argument, it must be concluded that the wave period remain unchanged<br />

while the wave pass from the open water to the ice-covered water. In order to satisfy<br />

the basic equation:<br />

A<br />

i = [10]<br />

the wave length must be modified as required by the phase velocity relation for ice­<br />

covered water, according to Eq. [8].<br />

Figure 2 shows how Ai varies with wave period for various ice thicknesses and<br />

water depths.<br />

[]]<br />

[8]<br />

[9]<br />

845


Experimental data have shown that the average ice thickness on Lake St. Peter<br />

can be related to a semi-empirical formula (Carter 1977):<br />

h = 0,024 1 1 / 2 [28J<br />

From the very few available data, Carter and Ouellet (1980) have estimated that,<br />

for a bulk carrier type ship sailing in the middle of the navigation channel, the<br />

average wave height reaching the bordering ice cover can be evaluated by:<br />

H = 0,037 V<br />

for ship speed less than 12 knots.<br />

Figure 6 shows the allowable speed predicted for "lakers" which generate the<br />

highest waves for a given velocity.<br />

Conclusions<br />

The application of the theoretical results to Lake St. Peter using average values<br />

seems to yield promising insights into a better understanding of the <strong>com</strong>plex problem<br />

of the erosion of border ice under the action of waves from passing ships.<br />

However, even though the results achieved by the application of the proposed<br />

theory are very optimistic, it is well recognized that a <strong>com</strong>parison with experimental<br />

data is necessary to properly assess its validity.<br />

Acknowledgements<br />

The authors wish to thank Transport Canada for permission to present this<br />

publication. Advice and assistance from Mr. N.E. Eryuzlu, Transport Canada. Waterways<br />

Development Division, are also greatly acknowledged, Part of the materials of this<br />

paper is obtained from our study financed by D.S,S, and Transport Canada.<br />

Notations<br />

The subscripts 0 and i refer respectively to open and ice-covered waters.<br />

C Phase velocity (m/s)<br />

E Elastic modulus of ice (N/m 2 )<br />

EI Average kinetic energy of water per unit area (N-m/m2)<br />

E2 Average potential energy of water per unit area (N-m/m2)<br />

E3 Average kinetic energy of ice per unit area (N-m/m2)<br />

E4 Average potential energy of ice per unit area (N-m/m2)<br />

H Wave height (m)<br />

1 Freezing index (OC-days)<br />

M Flexural . . d . t f 1 t h 3 E<br />

r1g1 1 y 0 a p a e: 12 (1-v2) (N-m)<br />

T Wave period (s)<br />

[29J<br />

849


U<br />

V<br />

a<br />

d<br />

g<br />

h<br />

k<br />

P<br />

s<br />

t<br />

x, y, z<br />

n<br />

A<br />

\)<br />

P<br />

Pi<br />

a<br />

w<br />

<br />

References<br />

Group velocity (m/s)<br />

Ship speed (knots)<br />

Wave amplitude (m)<br />

Water depth (m)<br />

Gravitational acceleration (9,8 m/s2)<br />

Ice thcikness (m)<br />

Wave number 2 n/A (rad/m)<br />

Pressure at the ice-water interface (N/m2)<br />

Surface tension of water (0,074 N/m)<br />

Time (s)<br />

Cartesian coordinates<br />

Sinusoidal displacement of ice plate (m)<br />

Wave length (m)<br />

Poisson's ratio: 0,3<br />

Density of water (kg/m 3 )<br />

Density of ice (kg/m3)<br />

Stress in ice sheet (N/m 2 )<br />

Angular frequency, 2n/T, (rad/s)<br />

Velocity potential (m 2 /s).<br />

Carter, D. and Ouellet, Y. 1980 "Fracture of a Solid Ice Cover by Wind-Induced or<br />

Ship-Generated Waves". Report prepared for Ministry of Transport under contract<br />

lSV79-000ll, 90 pages.<br />

Carter, D. 1977 "Ice Thickness in the St. Lawrence Waterway" Report prepared<br />

for Ministry of Transport, File No.: 80l0-ll5lCGAA, 22 pages.<br />

Carter, D. 1971 "Lois et mecanismes de l'apparente fracture fragile de la glace<br />

de riviere et de lac" These de doctorat, Dep. de Genie civil, Section Mecanique des<br />

Glaces, Univ. Laval, Quebec.<br />

Ewing, M. and Crary, A.P. 1934 "Propagation of Elastic Waves in Ice" Physics,S, 2,<br />

pp. 181-184.<br />

Lamb, H. 1932 "Hydrodynamics" Cambridge University Press, 6th Edition 1975.<br />

Peters, A.S. 1950 "The Effect of a Floating Mat on Water Waves" Commun. Pure<br />

Appl. Math., Vol. 3, pp. 319-354.<br />

Robin, G. de Q. 1963 "Wave Propagation through Fields of Pack Ice" Phil. Trans.<br />

Roy. Soc. London, Ser. A, 255 (1057), pp. 313-339.<br />

850


Wadhams, A. 1973 "Attenuation of Swell by Sea Ice" Jour. Geophys. Res., Vol. 78,<br />

No. 78, pp. 3552-3561.<br />

Weitz, M. and Keller, J.B. 1950 "Reflection of Water Waves from Floating Ice in<br />

Water of Finite Depth" Commun. Pure Appl. Math. Vol. 3, pp. 305-318.<br />

851


B.D. Pratte and G.W. Timco<br />

Research Officers<br />

ABSTRACT<br />

A NEW MODEL BASIN FOR THE TESTING<br />

OF ICE-STRUCTURE INTERACTIONS<br />

Hydraul ies Laboratory<br />

National Research Council<br />

Ottawa<br />

Canada<br />

The Hydraulics Laboratory of the National Research Council of Canada has con­<br />

structed a facility for model testing the dynamic interaction between a structure and<br />

an ice cover. The facility consists of a refrigerated chamber (29 m long, 10 m wide<br />

and 1.2 m deep in which carbamide (urea) doped model ice is grown. The design of the<br />

test facility has many novel features including duet work along one side of the<br />

chamber which blows cold air at various speeds across the ice surface to hasten the<br />

ice formation, a moveable insulated curtain which separates the main tank from the<br />

rigging area, and full instrumentation of the air, ice and solution for monitoring<br />

the changes in temperature and thermal state during the seeding process, growth and<br />

warm-up of the ice sheet. In this paper the features and operational characteristics<br />

of this test facility are described.<br />

857


1.0 Introduction<br />

With the rapidly escalating activity In Canada's Ice covered waters, It is<br />

necessary that design data on the interactive forces and displacements between struc­<br />

tures and ice covers be obtained. Not only must ships and drilling rigs move through<br />

the Ice, but also fixed structures such as drilling platforms, artificial islands,<br />

docks and breakwaters, etc. must be able to withstand the forces of ice covers moving<br />

past them. To better understand the forces which a structure experiences by relative<br />

motion between it and an ice cover, a few laboratories have constructed modelling<br />

facilities to simulate to scale the ice-structure Interactions. Host of these facil­<br />

ities, however, are mainly concerned with the testing of ship models in both uniform<br />

and broken ice covers. This type of model testing has greatly advanced the science<br />

of designing efficient ice breaking hull forms. To date, however, only a handful of<br />

tests have been performed on measuring the forces which a moving Ice sheet can exert<br />

on a stationary structure. This, in spite of the increasing importance of such<br />

stationary platforms in Arctic regions. For ice covers moving past fixed or moored<br />

structures, the relative speed is considerably lower than for ships in ice, but the<br />

forces may be much greater due to the large size and blunt shape of many structures.<br />

Recently the Hydraulics laboratory of the National Research Council of Canada in<br />

Ottawa constructed a refrigerated cold room which will be used to study such<br />

ice-structure interactions on a model scale. In addition, this facility will be<br />

used to quantitatively document mechanical properties and growth characteristics of<br />

model ice so that more meaningful results can be obtained from tests on small model<br />

structures. Upon <strong>com</strong>pletion of the carriage in late 1981, basic research on forces<br />

and motions of structures moving relative to the ice cover will begin. Such research<br />

will <strong>com</strong>plement the state of knowledge and be used to assist <strong>com</strong>mercial ice tank<br />

laboratories in evaluating their own testing on proposed Arctic structures. This<br />

paper describes the features and operational characteristics of this test facility.<br />

2.0 General layout<br />

Fig. 1 shows the general layout of the facil ity. The cold room, which is built<br />

inside the existing laboratory, is 29 m long, 10 m wide and 4.9 m high. Its walls<br />

consist of 10 cm thick polyurethane foam insulation (R20) between enamelled steel<br />

exterior and white fibreglass reinforced plastic (FRP) interior wall panelling.<br />

The basin itself is 21 m long, 7 m wide and 1.2 m deep to the top of the walls.<br />

The tank is filled with a carbamide (urea) solution [1,2] to a depth of 1 m. Both<br />

the basin walls (0.4 m thick) and the basin floor (0.2 m thick) are constructed of<br />

heavily reinforced poured concrete (5000 psi (34.5 HPa) minimum <strong>com</strong>pressive strength}.<br />

The entire tank is sealed with a black trowel-on rubberized ashphalt coating TP90V,<br />

and then painted Polarcote white. Because several of the existing ice tanks have<br />

B58


Photo I View of Ice Basin from Northwest Corner showing<br />

Air Supply Outlets, Ceil ing Mounted Thermocouples,<br />

Service Carriage, Walkway, Insulated Curtain<br />

Stowed Along Right Wall, and Board Frozen into Ice<br />

where Curtain is Normally Drawn during the Freezing<br />

Process<br />

861


862<br />

Photo 2 Testing Ice Sheet Properties. View from Southwes t<br />

End Showing Ductwork, Service Carriage, Evaporators<br />

and Stowed Curtain


Photo 5 Clearing the Ice Sheet Using Scraper Screen<br />

on Service Carriage - Movement to Right<br />

Photo 6 Pushing the Old Ice Sheet over Ramp into the<br />

Melt Pit (Pump and Spray Bar at Right)<br />

865


After ensuring that an ice skin covers the whole surface of the solution, the fans<br />

are restarted and the ice sheet Is grown at an ambient air temperature of -21°C.<br />

Following the freeze, the refrigeration system is shut off and the room temperature is<br />

allowed to rise to O°C. This procedure warms the ice sheet and aids in reducing its<br />

strength for the model studies.<br />

After the testing with the ice sheet is <strong>com</strong>plete, the ice is removed from the<br />

tank in the following way. The air bubbler pipe near each side wall is started in<br />

order to free up the ice from the walls. The water from the melt pit is then pumped<br />

into the main tank until the water in the tank is overflowing back into the melt pit.<br />

Using the screen attached to the service carriage (Photo 5), the ice is manually<br />

pushed into the melt pit (Photo 6), where it is melted using the heater and spray bar<br />

system. Once the ice is melted, the pump and heater are shut off and the plumbing<br />

lines are drained.<br />

4.0 Future Work<br />

The main structures carriage will be operational this year. It will be propelled<br />

by two pinions driving onto racks mounted on each rail, and powered by a 40 HP d.c.<br />

motor. The carriage speeds will be 0 to 50 cm/s for structures testing, with higher<br />

speeds available if required. The carriage will be exceedingly stiff with a natural<br />

frequency above 25 Hz and the ability to exert up to 5 tons horizontal force. Load<br />

cells will be used to mount the structures to be tested to the carriage. The car­<br />

riage speed and all force data will be recorded either on-board or in the control room<br />

for analysis by <strong>com</strong>puter.<br />

Basic research on forces exerted by moving Ice sheets on various elementary<br />

structures will be carried out. Such structures include circular piles, sloping<br />

cones, vertical and sloping faces such as the shores of islands, breakwaters, etc.<br />

More <strong>com</strong>plicated designs will also be tested as they evolve for Arctic applications.<br />

5.0 References<br />

1. Timco, G.W., "The Mechanical and Morphological Properties of Doped Ice: A Search<br />

for a Better Structurally Simulated Ice for Model Test Basins", Proc. POAC '79,<br />

Vol. I, pp. 719-739, Trondheim, Norway, 1979.<br />

2. Ti meo, G. W., "A Compar i son of Severa 1 Chem i ca 11 y-Doped Types of Mode 1 Ice", Proc.<br />

IAHR Symp. on Ice (in press), Quebec City, Canada 1981.<br />

3. Sandell, D.A., "Urea Ice Growth in a Large Test Basin", Proc. IAHR Symp. on Ice<br />

866<br />

(in press), Quebec City, Canada, 1981.


W.B. Tucker III, Geologist<br />

and<br />

Research<br />

W.D. Hibler III, Physicist<br />

PRELIMINARY RESULTS OF ICE MODELING IN THE<br />

ABSTRACT<br />

EAST GREENLAND AREA<br />

U.S. Army Cold Regions<br />

Research and Engineering<br />

Laboratory, Hanover, NH<br />

A sea ice model which employs a viscous-plastic constitutive law has been applied<br />

to the East Greenland area. The model is run on a 40 km spatial scale at 1/4 day<br />

time steps for a 60-day period using forcing data beginning 1 October 1979. Preliminary<br />

results verify that the model predicts reasonable thicknesses and velocities<br />

well within the ice margin. Separate simulations show that thermodynamics only and<br />

free drift with thermodynamics produce inadequate results. In particular, the free<br />

drift simulation produces unrealistic ice trajectories with excessive drift toward<br />

the coast and unreasonable nearshore thicknesses. The net results of these simulations<br />

tend to verify that internal ice stress, thermodynamics, and ice import must<br />

be considered to properly model this region.<br />

INTRODUCTION<br />

The East Greenland area is unique to this hemisphere because it is an area of<br />

confluence of polar and temperate systems for both atmosphere and ocean. The <strong>com</strong>plex<br />

nature of air-sea interaction here is further <strong>com</strong>plicated by the presence of sea ice.<br />

This ice cover is highly variable, with large changes in the extent occuring both on<br />

a seasonal and interannual basis.<br />

Wadhams [11 has extensively reviewed the literature which deals with the sea ice<br />

in this area. It is well known that the major <strong>com</strong>ponents which govern the ice balance<br />

are the thermodynamic balance at the sea surface, the wind and water stresses upon<br />

the ice, the Coriolis force and the internal ice stress. Additionally, there is a<br />

flux of ice into the East Greenland area from the Arctic basin which contributes to<br />

the mass balance [21. The relative importance of these <strong>com</strong>ponents to the region has<br />

not been made clear. More recent studies [1,31 have identified smaller scale proces­<br />

ses such as wave induced pulverization and warm eddies, which may contribute to the<br />

ice edge location on a synoptic scale.<br />

To begin to sort out the role of some of these processes, a first order sea ice<br />

modeling study seemed relevant. Initially, we decided that it would be most appropriate<br />

to apply a model that has been used successfully in studies of the sea ice in the<br />

USA<br />

867


A 40 km, 31 x 45 grid was established for the simulations. The grid location is<br />

shown in Figure 1. As the model allows for free boundaries (which allow inflow and<br />

outflow), these are designated for the Fram Strait (north), the Denmark Strait (south)<br />

and the eastern boundary. In order to satisfy stability criteria, a 1/4 day (21600 s)<br />

time step was required. All other parameters were identical to those used by Hibler<br />

[4] in the Arctic basin study with exception of the Coriolis parameter and P*, a<br />

constant used in determining ice strength. Here, the Coriolis term was calculated<br />

for each grid point location. The constant p* was set to four times that used in<br />

the Arctic basin simulations (5.0·10 3 Nm- 1 ). This change was implemented when initial<br />

tests showed the ice velocities to be excessive, presumably due to the large magni­<br />

tudes and variability of the daily winds. The previous Arctic simulations [4] used<br />

8-day averaged winds which inherently provided spatially and temporally smoothed<br />

fields.<br />

Due to <strong>com</strong>puter time limitations at our facility, we were limited to a 60-day<br />

simulation period. We chose the period October through November, 1979 because it is<br />

a season of relatively rapid ice expansion and because position data for drifting<br />

buoys located on the ice were available for this time period [7]. For the initial ice<br />

field, <strong>com</strong>pactness was digitized from the published ice chart [8] for 2 October 1979.<br />

Thickness was estimated by allowing it to vary linearly with latitude, 1.0 m at 67°N<br />

to 3.2 m at 83°N. These estimates seemed reasonable based on data reported from<br />

submarine transects of the area [9,10].<br />

Four basic simulations were performed in this study. Three were designed to test<br />

the response of the model 1) to thermodynamics only (no ice dynamics), 2) to ice<br />

import through the Fram Strait, and 3) to ice interaction (the importance of the ice<br />

stress term). The final test was designed to evaluate the general performance of<br />

the <strong>com</strong>pleted model in this area. For the thermodynamics only run, all ice velocities<br />

were set to zero and the ice was allowed to grow and ablate according to the modified<br />

ice growth rates for the 60 day period. The response of the model to ice inflow from<br />

the Arctic basin was evaluated by not assigning constant thickness to the northern<br />

free boundary cells as was the normal procedure. Without this specification (called<br />

the no inflow case), the model calculates open cell thicknesses as an average of the<br />

thicknesses in the adjacent cells inside the boundary and will be forced to run out<br />

of ice if large southward advection occurs. (Thus, this is not a true no inflow<br />

test; we would have to specify zero thicknesses for the open cells. This is more a<br />

test of a true northern open boundary). In the ice interaction test (referred to as<br />

free drift case), the ice strength was set to zero, effectively damping out the in­<br />

ternal ice stress. The final simulation (the standard case) included specification<br />

of inflow cell thicknesses, ice strength, and thermodynamics.<br />

870


Current and Wind Fields.<br />

RESULTS AND DISCUSSION<br />

The 60-day averaged wind velocity and ocean current fields are shown in Figures<br />

2 and 3. The most significant feature of the wind field is that the narrow band of<br />

northerly winds follows the continent almost precisely, indicating the large influence<br />

of topography on the sea level pressure field. Also noteworthy is the fact that the<br />

winds immediately to the east of this high velocity stream are southerly over most of<br />

its length. On the other hand, the current field is quite smooth as would be expected<br />

from a temporally constant dynawic height field.<br />

Ice Thickness.<br />

The 2 October 1979 ice thickness field used to initiate the model runs is shown<br />

in Figure 4. All thickness fields represent the average ice thickness for each grid<br />

cell, that is the product of the actual ice thickness and the <strong>com</strong>pactness (0.0 tol.O).<br />

We chose the 0.25 m contour to represent the ice edge because thinner ice is subject<br />

Figure 2. 60-day averaged wind velocity<br />

field.<br />

Figure 3. Current velocity field.<br />

871


)<br />

Figure 4. Initial average thickness<br />

field.<br />

Figure 5. Average thickness field<br />

for the thermodynamics only simulation<br />

after 60 days. Dashed line<br />

is actual ice edge position for<br />

2 December 1979. Contour values<br />

are in meters.<br />

to large variations due to growth and ablation. Later studies will examine the<br />

variability and diffusivity of the ice edge in detail.<br />

Figure 5 shows the result of the 60 day thermodynamic simulation. Also shown' is<br />

the observed ice edge position for 2 December 1979 [8]. This simulation shows that<br />

thin ice «1.0 m) has extended to the south and east, but the thicker ice categories<br />

have changed very little. Previous simulations that excluded the oceanic heat flux<br />

showed that thin ice covered the entire grid. The net effect then is that the param­<br />

eterized oceanic heat flux is <strong>com</strong>pletely dominating the ice edge location in this<br />

simulation. It is also apparent that too much ice production is taking place in the<br />

Denmark Strait area (lower part of grid), possibly indicating a lack of proper ocean<br />

heat flux parameterization in that region unless advection is the mechanism responsi­<br />

ble for moving this ice closer to the coast. The fact that the growth rates for<br />

thicker ice are much smaller than for thin ice and open water is evidenced by the<br />

small change in the location of the thicker ice contours (>1.0 m).<br />

The effect of the ice dynamics on the thickness field can be appreciated from<br />

Figure 6 which shows the 60-day thickness field for the standard run. Several<br />

872


Figure 8. Average thickness field for<br />

free drift simulation.<br />

" .....<br />

Figure 10. 60-day average velocity<br />

field for standard simulation.<br />

874<br />

Figure 9. 60-day average velocity<br />

field for free drift simulation.<br />

which signifies that ice import from the<br />

Arctic basin is important to the region. A<br />

longer simulation will verify whether the ice<br />

to the south also be<strong>com</strong>es thinner as the ice<br />

there continues to advect out of the southern<br />

boundary.<br />

The final thickness plot, shown in Figure<br />

8, is the result of the free drift simulation.<br />

with zero strength, thus no ice interaction<br />

or resistance to deformation, the ice builds<br />

to physically unrealistic thicknesses along<br />

the coast. The necessity of allowing for ice<br />

interaction in this region for any modeling<br />

effort is clearly demonstrated by this figure.<br />

Free drift may be applicable on a localized<br />

scale for very short term forecasts but it is<br />

certainly not sufficient for the region as a<br />

whole over time periods on the order of months.


PACK ICE DRIFT AND WEATHER IMPACT<br />

A Pilot Study off East Greenland<br />

Ren€ Zorn Danish Hydraulic Institute Denmark<br />

Hans H. Valeur Danish Meteorological Institute Denmark<br />

ABSTRACT<br />

Several mathematical models on ice drift have been developed since N.N. Zubov published<br />

his formulae. Common for most of them have been the lack of sufficient verification<br />

data, since most ice observations are too insufficient to allow evaluation of short<br />

term fluctuations <strong>com</strong>pared with weather data. The aim of this paper was to validate<br />

the above mentioned drift ice theory. Unfortunately it turned out that after obtaining<br />

ice drift observations it was the weather data that were insufficient. Instead it is<br />

shown that ice drift information might provide wind information to verify weather<br />

charts.<br />

An environmental research programme in August and September 1980 for the benefit of<br />

possible future oil prospecting in and off East Greenland created a possibility to<br />

investigate the ice conditions off East Greenland between 6S o N and 78 0 N as related to<br />

weather conditions.<br />

Two drift bouys were deployed in the area, and 22 reconnaissance flights were executed.<br />

The data from those platforms supplied with data from satellite images were <strong>com</strong>pared<br />

with the actual weather maps and climatic current data.<br />

1. INTRODUCTION<br />

In connection with a Marine Geophysical Survey Prngramme and Environmen­<br />

tal Studies Offshore East Greenland, an operational ice reconnaissance<br />

and drift ice study was executed in 1980 for Geological Survey of Green­<br />

land (GGU) and Grenland Technical Organization (GTO) both under the<br />

Ministry for Greenland.<br />

Results from the above programmes have been used for preparation of the<br />

present paper.<br />

879


2. GENERAL<br />

During August and September 1980 ice reconnaissance flights were<br />

carried out 2 or 3 times a week along the East Coast of Greenland /1/,<br />

departing from Reykjavik, Iceland or Narssarssuaq, Greenland. During<br />

and after the flights ice charts were prepared and transmitted to the<br />

Danish Meteorological Institute (DMI) in Copenhagen, Denmark. DMI also<br />

made interpretations of satellite images (NOAA 6) and weather analysis<br />

and forecast charts. All results were transmitted daily from Denmark<br />

to the survey vessels offshore East Greenland via radio facsimile,<br />

Fig. 1.<br />

Fig. 1. Communication Diagram<br />

880


The Environmental Study Offshore East Greenland 1979 and 1980, /2/,<br />

which included a drift ice study was based on board a survey vessel<br />

which was mainly operating in the area offshore Scoresbysund and Me­<br />

stersvig.<br />

The object of these experiments was to determine drift ice patterns and<br />

velocities by means of an automatic NIMBUS-6 station placed on an ice<br />

flow (ice island).<br />

The drift study covered the period August 3 to August 10, 1979; and<br />

August 20, 1980, to February 16, 1981 and results are given in Section<br />

3. Selected ice observations have been analysed together with weather<br />

analyses for the same period to determine sea ice drift and the re­<br />

sults are given in Section 4.<br />

3. NIMBUS-6 DRIFT ICE STUDY<br />

On August 3, 1979, an automatic NIMBUS-6 station was placed on an ice<br />

floe at position 73 0 36'N/19 0 52'W. The dimensions of the ice floe (a<br />

floe berg?) were approximately 60 x 40 m with a total thickness of<br />

about 30 m.<br />

The station has been tracked until August 10, 1979, when the ice floe<br />

was grounded for half a year until the station was lost, Fig. 2. The<br />

-1<br />

mean drift speed has been calculated to about 0.13 m'sec in the pe-<br />

riod of drift.<br />

On August 20, 1980, a second automatic NIMBUS-6 station was placed on<br />

an ice floe (ice island) at position 71 0 55'N/21 o 33'W. The approximate<br />

dimensions of this floe were about 450 x 150 m, with a height of 6 m<br />

above the sea surface and a draft of about 20 m.<br />

The station has been tracked from August 20, 1980, to February 16, 1981,<br />

by satellite, and from August 20 to September 21, 1980, also by air­<br />

craft. The track is shown in Fig. 2.<br />

881


• SATELLITE OBSERVATIONS<br />

o AIRCRAFT OBSERVATIONS<br />

DAY I MONTH -YEAR<br />

'50 100IO;M<br />

f---+I------il<br />

Fig. 2 Drift of automatic NIMBUS-6 stations placed on ice floes off<br />

882<br />

East Greenland August 1979 and August 1980 to February 1981.


The drift of the ice island reflects the currents at 20 m depth which<br />

may well be different from the surface currents. The time lag between<br />

wind and current grows with depth, so ice island drift will not provide<br />

information of immediate wind conditions, but will show mean current<br />

conditions through some period of time.<br />

4. ANALYSIS OF ICE OBSERVATIONS<br />

The ice charts of August 18 and September 11, 1980 (Figs. 4, 5 and 6)<br />

show the typical distribution of sea ice during the period of investiga­<br />

tion. As usual for that time of the year open water is prevalent in the<br />

fjords and along the coast, while a 100-200 km wide belt of pack ice of<br />

various concentration is present outside. Figs. 5 and 6 depict the ice<br />

concentration on September 11, 1980 as derived from aerial reconnais­<br />

sance and satellite images, respectively. Generally the two sources<br />

agree reasonably well, yet discrepancies, especially regarding concen­<br />

tration indications, do occur now and then.<br />

Fig. 4 Satellite image, August 18, 1980.<br />

884<br />

(Numbers indicate ice con­<br />

centrations in tenths,<br />

dotted areas indicate less<br />

than 1/10 ice concentration,<br />

heavily shaded areas indica­<br />

te 10/10 ice concentration).


It is a well known fact that the ice drift is determined by wind (di­<br />

rectly and through wind generated currents) '- with a time lag of about<br />

one day - and by the gradient and tidal currents. Therefore, under<br />

calm weather conditions (Fig. 7) the drift pattern should be simple<br />

with a weak negative vorticity, determined by the East Greenland Current<br />

as was the case in the interval August 24-25, 1980 (Fig. 8) where the<br />

drift was towards the S and the SSE, the speed varying from 0.06 m.sec- l<br />

near the coast to more than 0.32 m.sec- l close to the ice edge.<br />

Fig. 7 Weather Chart, August<br />

25, 1980 1200 GMT<br />

Fig. 8 Ice drift, August 24-25,<br />

1980.<br />

(Open arrows indicate movement of<br />

ice edge, while solid arrows indicate<br />

drift of individual floes)<br />

On the other hand, wind opposing the gradient current will reduce the<br />

drift (e.g. August 25-27, 1980 (Fig. 9) when the winds through the pre­<br />

ceding day and night, Fig. 3, were from the Sand SW thus causing a<br />

drift towards the East), while downcurrent winds will increase the drift<br />

velocity (e.g. the southernmost current vector on Fig. 11, August 28-29,<br />

1980) .<br />

886


Fig. 9 Ice drift, August 25-27,<br />

1980.<br />

\ - . -<br />

iJ :<br />

Fig. 10 Ice drift, August 27-28,<br />

1980.<br />

In a region like the present where weather data is utterly sparse, the<br />

weather analysis is as a rule extremely uncertain, whence ice drift<br />

vectors could be of considerable value in adjusting wind calculations.<br />

However, to avoid hidden subjectivity in stating the impact of wind on<br />

the ice drift, the weather charts in the present study have deliberately<br />

been analysed by a very experienced analyst (Class II meteorologist<br />

mr. Leif Rasmussen) not knowing the ice drift.<br />

The drift August 27-28 (Fig. 10) may serve as an example of how the ana­<br />

lysis may be adjusted: North of 76 0 30'N the drift was generally towards<br />

the E and the NE, which might indicate that the calculated northerly<br />

wind of 10 kts on August 27 at position 77 0 20'N/13 0 W (Fig. 3) actually<br />

should be 20 kts from the South.<br />

-<br />

......<br />

887


On the other hand the calculated 20 kts wind from the south at position<br />

7S o N/13 0 W may actually have been zero or weak northerly, causing a drift<br />

speed of nearly 0.43 m·sec- l around this position. A displacement of the<br />

indicated high pressure of about 150 nm towards the NNE on the weather<br />

map would fit these winds. The drift pattern August 28-29 (Fig. 11) is<br />

similar to that of August 27-28, 1980.<br />

Fig. 11 Ice drift, August 28-29, 1980.<br />

The other period selected for drift investigation September 8-12, also<br />

shows great local variations and seems to confirm the feasibility of<br />

using ice floes as a tool to adjust the weather analysis.<br />

On September 8-9, 1980 (Fig. 12) the drift was weak and diverging around<br />

7SoN. North of this latitude the drift seems to have been slightly to­<br />

wards the North in spite of indicated strong wind on September 7, 1980<br />

from northerly directions (Fig. 3). This may go to show that the pres­<br />

sure difference indicated by the isobars in the Greenland Sea area on<br />

September 7 and 9 (Fig. 13) was actually concentrated in the eastern<br />

8M


part. Hence, the isobars should be moved eastwards indicating a weak<br />

gradient at the positions indicated in Fig. 3 with calm weather and<br />

later southerly winds.<br />

Fig. 12 Ice drift, September 8-9,<br />

1980.<br />

Fig. 13 Weather chart, September<br />

7, 1980.<br />

September 9-11 and 11-12, 1980 (Figs. 14 and 15) showed a marked nega­<br />

tive rotation with considerable local variations. Only south of 75N the<br />

drift seems to have had a southerly direction (up to 0.32 m.sec-lat the<br />

edge), while the tendency north of this latitude was more a drift to-·<br />

wards the East.<br />

889


Fig. 14 Ice drift, September<br />

5. CONCLUSION<br />

9-11, 1980.<br />

Fig. 15 Ice drift, September<br />

11-12, 1980.<br />

The velocity values as derived from the present drift ice study agree<br />

fairly well with those found by Malmberg et al. /3/. In the inner shelf<br />

region a general decreasing current <strong>com</strong>ponent with depth is a <strong>com</strong>mon<br />

feature, with maximum currents of about 0.35 m'sec- l at the surface.<br />

The weather data were too sparse to allow any test of drift models to be<br />

made in this pilot study. However, the study did show the <strong>com</strong>plexity of<br />

the drift pattern, - similar to that found further to the south, by<br />

Malmberg et al., but much more irregular than found by Vinje /4/.<br />

Further it appears that the drift of ice floes may serve to verify historical<br />

weather charts thus proving a tool to improve the weather analysis.<br />

890


6. REFERENCES<br />

/1/ GGU, Ice Reconnaissance along the East Coast of Greenland, 1980.<br />

December 1980.<br />

/2/ GTO, Environmental Studies Offshore East Greenland, 1980. April<br />

1980.<br />

/3/ S.-A. Malmberg, H.G. Gade and H.E. Sweers, Current Velocities and<br />

Volume Transports in the East Greenland Current of Cap Nordenskjold<br />

in August-September 1965, Reykjavik 1972.<br />

/4/ Torgny E. Vinje, Sea Ice Studies in the Spitsbergen-Greenland Area.<br />

Oslo 1977.<br />

891


t. K<br />

P = p*....!. N<br />

to<br />

in which p* and K are empirical constants and to is a reference ice thickness. p* is<br />

presumably related to the yield strength of the macroscopic ice medium. The pressure<br />

modifier, N K , serves to reduce the ice pressure substantially when the ice area concentration<br />

is less than one.<br />

Model Calibration<br />

The model was applied to Lake Erie for the period January 15, 1979 to January 17,<br />

1979 ( a period of 49hours) during which significant freezing and movement of the ice<br />

field had been observed to occur. Side-looking airborne radar images (SLAR) and the<br />

ac<strong>com</strong>panying ice chart products produced by the U.S. Coast Guard Ice Navigation Center<br />

in Cleveland, Ohio were the principal observational data available for model calibration<br />

(see Figures 1 and 2).<br />

During this freezing period the magnitude of Em was <strong>com</strong>puted differently for<br />

ice covered areas than for open water areas. For ice covered areas, Em was represented<br />

as follows,<br />

where Ta = local air temperature and a = an empirical coefficient.<br />

areas, Em was given by,<br />

(5)<br />

(6)<br />

For open water<br />

where a = an empirical coefficient. The wind speed and the air temperature at any<br />

position over the lake were obtained by weighting the observed values at n meteorological<br />

stations (in this case, Toledo, Cleveland, and Buffalo) using a weighting procedure<br />

[11) and correcting overload wind measurements to overlake values [13),<br />

The following values of the constants were utilized in this calibration effort:<br />

p* = 5000 N/m2; to = 0.5 m; K = 20; e = 2.0; k 2 (max) = 10 10 Ns/m 2 ; k 2 (min) = 104 Ns/m 2 .<br />

e denotes the ratio of the lengths of the principal axes of the assumed two-dimensional<br />

stress ellipse for the ice medium [9). Only the constants a and a from equations<br />

(6) and (7) were varied in this calibration effort: case (1) [a = 4x10- 6 , a = 2x10- 5 );<br />

case (2) [a = 3x10- 6 , a = 6X10- 5 ); case (3) [a = 3x10- 6 , a = 3x10- 4 ). The water<br />

currents were assumed to be zero throughout the basin and the no-slip condition was<br />

applied to the ice velocity <strong>com</strong>ponent tangential to the lake boundaries. However,<br />

the <strong>com</strong>ponent of the ice velocity normal to an impenetrable boundary was expressed<br />

896<br />

(7)


differently depending on the direction of ice motion at the boundary. On a boundary<br />

from which ice mass moves lakeward (upwind boundary), the boundary condition requires<br />

that the gradient of the normal velocity <strong>com</strong>ponent be zero, while at a downwind<br />

boundary, the normal ice velocity <strong>com</strong>ponent must be zero. The initial conditions for<br />

N(x,y,o) and t.(x,y,o) were obtained from the SLAR image and ac<strong>com</strong>panying ice chart<br />

1 +<br />

product for January 15, 1979 (see Figure 1). The meteorological data, Va and Ta'<br />

were corrected every 3 hours according to the input of observed values at Toledo,<br />

Cleveland, and Buffalo. The initial ice velocities, Vi(x,y,o), were set equal to<br />

zero since the transient period for ice mass acceleration had been previously found<br />

to be on the order of 15 minutes [16].<br />

The model output after 49 hours is shown in Figures 3, 4, and 5. These figures<br />

portray the ice condition for January 17, 1979 in terms of ice drift velocities, ice<br />

area concentration, ice thickness, and ice pressure. The ice pressure field is<br />

presented as the product of the local <strong>com</strong>puted pressure and the local ice thickness<br />

(pt i ) which is representative of the hull pressure to be encountered by vessels transiting<br />

the ice field. Since the ice pressure has been parameterized in the present<br />

model and exact values are unknown, the plots of pt i are only qualitative. However,<br />

in this manner, it is possible to identify regions of high value of pt i which may be<br />

useful in charting vessel tracks through the ice field. The <strong>com</strong>puted ice condition<br />

for January 17, 1979 is in reasonable agreement with the observed conditions as portrayed<br />

by the SLAR image and ac<strong>com</strong>panying ice chart product for that same day (see<br />

Figure 2). Only the <strong>com</strong>puted values of Nand ti can be <strong>com</strong>pared with the data from<br />

Figure 2.<br />

Summary<br />

A fairly elaborate and versatile numerical model has been developed for use in<br />

forecasting ice conditions in the Great Lakes including a graphical output-retrieval<br />

scheme. The application of the model has been demonstrated by calibration against an<br />

observed ice transport episode in Lake Erie during January, 1979. A second calibration<br />

is being conducted for an observed ice tranport episode in Lake Erie (March, 1979)<br />

during which significant ice melting occurred. The heat exchange occurring in the<br />

prototype could be dealt with more explicitly [19], but the <strong>com</strong>putational effort and<br />

data requirements would be substantially greater.<br />

Further calibration efforts to selected portions of one of the Great Lakes is<br />

re<strong>com</strong>mended. Better observational data are needed for <strong>com</strong>parison with model output<br />

in order to make further adjustments in model coefficients and to reveal more fully<br />

the use of the model as a forecasting tool.<br />

897


[I)<br />

[2)<br />

[3)<br />

[4)<br />

[5)<br />

[6)<br />

[7)<br />

[8)<br />

[9)<br />

[10)<br />

[11)<br />

[12)<br />

[13)<br />

[14)<br />

[15)<br />

[16)<br />

[17)<br />

[181<br />

[19)<br />

898<br />

REFERENCES<br />

Campbell, W.J., "The Wind-Driven Circulation of Ice and Water in a Polar Ocean",<br />

J. Geophys. Res., Vol. 70, 1965, pp. 3279-3301.<br />

Campbell, W.J., and Rasmussen, L.A., "A Numerical Model for Sea Ice Dynamics<br />

Incorporating Three Alternative Ice Constitutive Laws", Sea Ice, Proc. of an<br />

Internat. Conf. Reykjavik, Iceland, May, 1971, pp. 176-189.<br />

Coon, M.D., "Mechanical Behavicr of Compacted Arctic Ice Floes", J. Petro. Tech.<br />

Vol. 26, 1974, pp. 466-470.<br />

Coon, M.D., Maykut, G.A., Pritchard, D.S., and Thorndike, A.S., "Modeling the<br />

Pack Ice as an Elastic Plastic Material", AIDJEX Bulletin, No. 24, 1974, pp.<br />

1-103.<br />

Doronin, Y.P., "On a Method of Calculating the Compactness and Drift of Ice<br />

Floes", AIDJEX Bulletin, No.3, 1970, pp. 22-39.<br />

Glen, J.W., "Thoughts on a Viscous Model for Sea Ice", AIDJEX Bulletin, No.2,<br />

1970, pp. 18-27.<br />

Hibler, W.O., "Differential Sea Ice Drift II: Comparison of Mesoscale Strain<br />

Measurements to Linear Drift Theory Predictions", J. Glaciology, Vol. 13,<br />

No. 69, 1974, pp. 457-471.<br />

Hibler, W.O., "A Viscous Sea Ice Law as a Stocastic Average of Plasticity",<br />

J. Geophys. Res., Vol. 82, 1977, pp. 3932-3938.<br />

Hibler, W.D .. "Modeling Pack Ice as a Viscous-Plastic Continuum: Some Preliminary<br />

Results", Proc. of a Symp. on Sea Ice Processes and Models, Unv. of<br />

Washington, Seattle, Wash., 1977, pp. 46-55.<br />

Neralla, V.R., and Liu, W.S., "A Simple Model to Calculate the Compactness of<br />

Ice Floes", J. Glaciology, Vol. 24, No. 90, 1979.<br />

Platzman, G.W., "The Dynamic Prediction of Wind Tides on Lake Erie", Meteorological<br />

Monographs, Vol. 4, No. 26:1-44, 1963.<br />

Pritchard, R.S., "An Elastic-Plastic Constitutive Law for Sea Ice", J. Applied<br />

Mech., Vol. 42, No.2, 1975, pp. 379-384.<br />

Resio, D. T., and Vincent, C.L., "Estimation of Winds over the Great Lakes",<br />

Miscellaneous paper, H-76-12, U.S. Army Corps of Engineers, Vicksburg, Mississippi,<br />

1976.<br />

Rothrock, D.A., "The Mechanical Behavior of Pack Ice", Annual Review of Earth<br />

and Planetary Sciences, Vol. 80, No.3, 1975, pp. 317-342.<br />

Rumer, R.R., Crissman, R.D., and Wake, A., "Ice Transport in Great Lakes",<br />

Contract No. 03-78-B01-104, Great Lakes Environmental Research Lab., NOAA, Ann<br />

Arbor, Mich., Sept. 1980.<br />

Rumer, R.R., Wake, A., Chieh, S-H, Fukumori, E., and Tang, G., "Internal Resistance<br />

of Lake Ice", Contract No. NA79RAC00124, Great Lakes Environmental Research<br />

Lab. NOAA, Ann Arbor, Mich., Sept. 1981.<br />

Udin, I., and Ullerstig, A., "A Numerical Model for Forecasting the Ice Motion<br />

in the Bay and Sea of Bothnia", Research Report No. 18, Winter Navigation<br />

Board, Norrkoping, Sweden, 1976.<br />

U.S. Coast Guard, "Lake Erie Ice Charts", U.S. Coast Guard Ice Navigation<br />

Center, Cleveland, Ohio, 1979-1980.<br />

Wake A., and Rumer, R.R., "Modeling the Ice Regime of Lake Erie", Proc. ASCE.<br />

Vol. 105, No. HY7, 1979, pp. 827-844.<br />

ACKNOWLEDGMENT<br />

This work is supported by the Great Lakes Environmental<br />

Research Laboratory, National Oceanic and Atmospheric Administration,<br />

U.S. Dept. of Commerce, Ann Arbor, Michigan.


900<br />

Fiqure 2. Lake Erie Ice Condition on January 17, 1979 (t=49.Z hrs)<br />

Methodology for Computing Nand ti from Ice Charts<br />

Code from Ice Chart: alaZa3a 4<br />

nln Zn 3<br />

a<br />

1<br />

a<br />

Z<br />

a<br />

3<br />

a<br />

4<br />

tenths of area with ice thickness, t<br />

1<br />

=0.5m<br />

tenths of area with ice thickness, t<br />

Z<br />

=0.ZZ5m<br />

tenths of area with ice thickness, t<br />

3<br />

=0 . 075m<br />

tenths of area with ice thickness, t<br />

4<br />

=0.OZm<br />

1<br />

N(x,y) = To E a n<br />

ti(x,y) = (E antn)/( E an)


THE BIOLOGICALLY IMPORTANT AREAS IN THE<br />

ARCTIC OCEAN<br />

Erkki Palosuo, prof.emer. University of Helsinki Finland<br />

ABSTRACT<br />

The shelf between Spitzbergen and Franz Josef Land is considered as<br />

a biologically important area in the open sea, and the Mackenzie<br />

Estuary and the Stefansson Sound are considered important in inshore<br />

areas. A general view of the biological productivity and its signifi-<br />

cance to the Arctic Basin ecosystem is given using these examples.<br />

1. INTRODUCTION<br />

Several scientists who have been working in the Arctic have noted how<br />

certain animals gather in particular places, known, for example, as<br />

hunting areas. Those who have had the opportunity to study the pri-<br />

mary biolog¥ and the rather simple food chains in the sea have<br />

realized that the Arctic Ocean is a fascinating environment, which<br />

has only just begun to be understood. Recent expeditions have shown<br />

that some particular areas are very important to the ecosystem of the<br />

Arctic.<br />

Of these "cradles of life", I shall mention the following:<br />

- the shelf between Spitzberegen and Franz Josef Land, a center of<br />

902<br />

the local polar bear population (Fig.l),


ut scare in the open sea.<br />

It is well known that the majority of nutrients reach the Arctic<br />

Ocean from Siberia, but the author has no information concerning the<br />

amounts involved. In the Stefansson Sound and its nearshore zone,<br />

the energy budgets for production indicate that about half of the<br />

carbon input is from terrestial sources, which include shoreline<br />

erosion and transport of humus detritus by rivers (13). It is impor­<br />

tant to note that the humus represents a large fraction of the ener­<br />

gy requirements of nearshore invertebrates. Its importance can be<br />

seen in the low 14C content of organisms during winter, when the<br />

flow of humus into sea is reduced, (report of prof. Donald M. Schell).<br />

In the summer, convection brings the nutrients from shallow bottom<br />

up to the upper water layer without any upwelling or other phenomena.<br />

The turbidity following this convection can considerably reduce the<br />

primary production.<br />

There is very little information available on the effect of a river<br />

on the sea. In the Mackenzie Estuary, the influence of the river<br />

seems to reach some 10 km out from the outer banks, where the gyra­<br />

tory circulation in the Beaufort Sea is dominant (8). The same situ­<br />

ation occurs in the Baltic Sea, too.<br />

We can conclude that estuaries and the nearshore zone from many<br />

distinct areas, in which the primary production and animal life<br />

depend on local condi tions·.<br />

3. THE CENTRAL PART OF THE ARCTIC BASIN.<br />

3.1. Nutrients<br />

The quantity of nutrients in the central part of the Arctic Basin<br />

varies according to the way they have entered the basin. As mention­<br />

ed above, most nutrients originate inland and are carried by rivers<br />

to the sea. Water from adjacent seas also enters the Basin. The<br />

Atlantic Gulf stream flows to the west of Spitzbergen and the sub­<br />

merges under the less saline Arctic water along the continental<br />

shelf to the north of the Siberian, Alaskan and Canadian Arctic (16b).<br />

Uppwelling finally raises this relatively warm watermass to the sur­<br />

face. Water from the Bering Sea eneters the Chukchie and Beaufort Seas.<br />

905


tion under the ice usually reaches a depth of 10 to 20 m. Such a<br />

shallow depth is favourable to the development of a bloom and thus<br />

to the whole primary productivity.<br />

In the shelf-break area of the Bering Sea, Dr. Alexander made the<br />

very significant observation that the necessary stratification and<br />

chlorophyll concentration occurred off the edge of the ice up to<br />

25 km seawards. The effect of melting ice on the stratification is<br />

much higher than the warming of water in summer or upwelling at the<br />

ice edge caused by wind-driven off-ice Ekman transport.<br />

There have been large annual variations in the seasonal extent of<br />

the ice edge in the Bering Sea (1). The same has also been observed<br />

in the other parts of the Arctic Ocean. The relationship between<br />

the spatial and temporal position of the seasonal ice edge might be<br />

very important in determing wheter sufficient nutrients are available<br />

to support the summer bloom, and in determing the rate of increase<br />

in the phytoplankton <strong>com</strong>munity.<br />

3.3. The role of ice algae<br />

An extraordinary phenomen occurs in the Arctic and in the Antarctic<br />

seas: the blooming of ice algae in and under the ice in the early<br />

summer. As shown by Meguro et.al. (11) the diatoms reproduce in brine<br />

solutions trapped in the microfissures between fine crystals of sea<br />

ice. they form a brownish layer near the bottom of ice.<br />

The main poin! of interest is firstly that they develop during the<br />

summer. The light intensity in the ice is more favourable for photo­<br />

synthesis than in the water under the ice. Meguro found no diatom<br />

activity at a distance greater than 0.3 m from the bottom of the ice.<br />

He postulated that diatoms are frozen into sea ice as it is formed<br />

and that they then grow in the early summer as a result of increases<br />

in nutrient and light.<br />

Ice algae bloom in the first year ice only. The surface warming of<br />

the ice causes brine to descend from the upper layers and none solar<br />

radiation reaches the second year ice below the older ice. Thus no<br />

ice flora will develope on the under side of the second year ice.<br />

Selective absobtion increase the temperature and therefore the<br />

porosity of the layer until the ice disintegrates. Second year ice<br />

907


later forms at the base of the ice sheet and there is no evidence<br />

908<br />

of the summer plankton layer being transferred into the second year<br />

ice. The chlorophyll content in the brown layers is from 40<br />

to more than 100 times greater than that in the sea water below the<br />

ice (10,16). An explanation to this is probably a considerable ex­<br />

change of salt and nutrients between the lower layers of ice in<br />

which the algae occur and the seawater beneath it (9).<br />

Later, when the ice has disappeared, a general phytoplankton primary<br />

production can take place. According to the observations of English<br />

(6) on drift station "Alpha", in 1956-58 in the high Arctic, phyto­<br />

plankton appear in the water for only a short time in summer. The<br />

chlorophyll concentrations suddenly increased at the end of June,<br />

but had already decreased in September. The average daily production<br />

for the summer was 5 to 6 mg c/m 2 d (14). Compared with values from<br />

other oceans, especially with Atlantic, this is rather low.<br />

The average annual production in the nearshore zones is 20 g C/m 2 a<br />

(7), and less in the open sea. However, the high production of ice<br />

algae means that the total primary production of the Arctic Ocean<br />

is not necessarily the smallest of all the oceans.<br />

4. REQUIREMENTS FOR A "CRADLE OF LIFE"<br />

The geographical areas in which the primary productivity is high are<br />

often marked by an abundance of mammals or other animals at the top<br />

of the food chain. The simplest food chain is, e.g. when the phyto­<br />

plankton is consumed by zooplankton which again is eaten by a bow­<br />

head whale. Slightly more <strong>com</strong>plicated food chains arise when small<br />

fish and Arctic cod consume the zooplankton and then seals and beluga<br />

whales devour the fish.<br />

The king of the Arctic, the polar bear, is at the top of the food<br />

chain. He catches seals from ice floes. A rich primary production in<br />

the upper water layer, a shallow sea area for rich bottom fauna, an<br />

abundance of seals and ice floes are needed to maintain the polar<br />

bear population. One of the areas is the shelf between Spitzbergen<br />

and Franz Josef Land (Fig.l). Its western part is only 200 to 300 m<br />

deep. A continous flow of drift is through the sounds between the<br />

islands renews the ice field. The drifting floes are mostly first<br />

year ice with many ridges (Figs 4 to 5). The original thickness of


910<br />

the sheet ice was 1.5 m, but in July it had melted to a thickness<br />

of 1 m. It is natural that Kongsoya and the two other islands belong­<br />

ing to the Kong Karls Land in the western part of the area form the<br />

mating place of the polar bear population. Polar bear cubs are born<br />

and raised up on these big islands (16).<br />

5. ENVIRONMENTAL CONCERN OVER POTENTIAL OIL SPILLS<br />

Protection of the Arctic Ocean and the Arctic in general must start<br />

from the fact that the Arctic ecosystem is relatively simple, with<br />

a small number of species and few feeding links. As Dunbar (4) has<br />

mentioned, the simplicity of the food chain has important effects<br />

upon the stability and vulnerability of populations. It has been<br />

found that Arctic populations fluctuate widely, and has been argued<br />

that simple ecosystems are more vulnerable to change than <strong>com</strong>plex<br />

ecosystems in temperate regions. The slow growth rate in Arctic<br />

ecosystem may well be dependent upon both food and temperature. This<br />

means that to animal and plant populations takes a long time to<br />

repair.<br />

Special attention must be paid to the areas which are conductive to<br />

biological production. It is obvious that the influence of these<br />

areas on the Arctic life is not limited to the areas themselves.<br />

They also provide energy for other regions.<br />

With regard to oil spillage, we must consider the microbes as de<strong>com</strong>­<br />

posers of oil. One surprising fact was found during "Ymer-80" expedi­<br />

tion: despite the low temperature of the water, the number of bacte­<br />

ria was e.S high as in the seas on middle latitude, even at great<br />

depths (16e), and were active at all depths. On other hand, de<strong>com</strong>po­<br />

sition by microbes can be slow, and the oil can remain for a long<br />

time in an ice covered area. If we allow oil transport during the<br />

short period of fundamental energy incorporation by the Arctic eco­<br />

system, an oil spill might be a catastrophe for the whole Arctic<br />

Ocean. In first place, it would be disastrous to the basic level of<br />

the ecosystem; its external effects on birds and mammals wouls take<br />

second place


REFERENCES<br />

1. Alexander, Vera and N.J. Niebauer. 1981. Limnology and Oceano­<br />

graphy. (In press.)<br />

2. Andersson, L. and D. Dyrssen. 1980. Report on the chemistry of<br />

seawater, XXIV. Department of analytical and marinen chemistry<br />

Chalmers University of Technology and university of GOteborg.<br />

3. Contributions to the Swedish Arctic Expedition "Ymer-80".<br />

(Report of Thor Larsen)<br />

4. Dunbar, M.J •• 1971. MCGill-Queen's University Press.<br />

5. Dunton,K. and Susan Schonberg. 1979. Environmental Assesment of<br />

the Alaskan Continental Shelf. Annual Reoprt (A.C.Broad). Nat.<br />

Oceanic Atmos. Admin., Boulder, Co.<br />

6. English,T.S •• 1961. AINA, Sci.Rep. 15.<br />

7. Fogg,G.E .• 1977. Phil. Trans. R. Soc. London, B 279.<br />

8. Fraker.M.A., C.D.Gordon, J. McDonald, J.K.B.Ford and G.Cambers.<br />

1979. Fisheries and Marine Service, Techn. Rep. 863.<br />

9. Martin, S •• 1970. Geophysical Fluid Dynamics, 1: 143-160.<br />

10. Maykut,G.A. and N.Untersteiner. 1971. Journ. Geophys., Vol.76:<br />

1550-75.<br />

11. Meguro H., I.Kuniyuki and H.Fukushima. 1967. Arctic, Vol.20:114<br />

12. Niemi,A. 1973. Acta Botanica Fennica, 100.<br />

13.Schell,D.M.,1980. Beaufort Sea Winter watch, Spec. Bull. 29:25-30.<br />

14. Strickland,J.D.H •• 1960. Fish.Res.Bd. Canada, Bull. 122.<br />

15. Sverdrup,H.U •. 1953. Joun.Cons.lnt.Explor.Mer. 18:287-295.<br />

16. Ymer 1981, Arsbok. (Expedition "Ymer-80"in Swedish).<br />

a.Palosuo,E. (Ice around "Ymer"): 46-50,<br />

b. Aagard,K., A.Foldvik and B.Rudels. (Physical oceanography):<br />

110-121,<br />

c. Dyrssen,D. (Chemie of the Arctic Ocean): 122-130,<br />

d. Hernroth,L. and L.Edler. (Planktonologically studies): 155-158,<br />

e. Kjellberg,S. (Microbiologically studies): 159-162.<br />

911


Austin Kovacs<br />

Rexford M. Morey<br />

Donald F. Cundy<br />

Gary Decoff<br />

POOLING OF OIL UNDER SEA ICE<br />

U.S.A. Cold Regions Res. and Eng. Lab<br />

Morey Research Co., Inc.<br />

U.S. Coast Guard Res. & Development Center<br />

U.S.A. Cold Regions Res. and Eng. Lab<br />

Abstract<br />

Ice thickness profiles were constructed for six fast ice locations in the V1C1nity<br />

of Prudhoe Bay, Alaska, using a radar echo sounding system. The sounding data revealed<br />

in detail the undulating relief of the bottom of the sea ice in which oil could<br />

pool up if released under the ice. In general, ice bottom morphology was found to reflect<br />

variation of the surface snow cover thickness and ice deformation. However, at<br />

several sites the ice bottom relief could not be correlated with these factors. Slush<br />

ice accumulations of up to 0.5 m were apparently the cause of this bottom roughness.<br />

Estimates of the volume of oil that could pool up in the ice bottom relief range from<br />

20,000 to 60,000 m3 /km 2• For undeformed fast ice with no bottom slush ice growth the<br />

potential pooling capacity varied from about 10,000 to 35,000 m3 /km 2• The effect of<br />

slush ice relief and structure on potential under-ice oil pooling is for the most part<br />

unknown.<br />

Introduction<br />

Offshore oil and gas exploration has been underway for some 10 years in the<br />

southern Beaufort Sea. Seismic studies and drilling results indicate significant oil<br />

and gas fields, and one can anticipate that within 5 years offshore resource production<br />

will be underway, with the attendant risk of oil being released under the sea<br />

ice. The first production will probably take place in areas that are seasonally<br />

covered by relatively undeformed fast ice.<br />

Information on under-ice relief is extremely limited and not well documented.<br />

Most ice thickness information consists of spot measurements, which are of no value in<br />

assessing under-ice oil pooling potential. Even where profile information exists it<br />

generally <strong>com</strong>prises ice thickness measurements that were taken along a single traverse<br />

of limited length and is therefore of uncertain value. Holtsmark [6), Roots [14) and<br />

Barnes et al. [1) showed the effect of snow cover thickness variation on the growth of<br />

sea ice. Hanson [4) described the accumulation and variation in snow cover thickness<br />

on sea ice during the course of a year. Hibler [5) reviewed the thermal processes responsible<br />

for sea ice growth and decay and discussed the <strong>com</strong>plex effects of varying<br />

snow cover thickness on the ice growth and decay process. His review and the aforementioned<br />

field studies clearly show that because winter snow cover acts as an<br />

insulator, reducing heat exchange from the sea water through the sea ice to the atmosphere,<br />

sea ice growth and therefore thickness vary inversely with snow cover depth<br />

variation. As a result the underside of undeformed annual sea ice has a rolling, hummocky<br />

topography that tends to mirror the long-term snow accumulation patterns on the<br />

ice surface.<br />

912<br />

USA<br />

USA<br />

USA<br />

USA


Ice thickness profile data were collected using a single antenna impulse radar<br />

sounding system. This system was similar to the one used by Kovacs [7] in 1976 to<br />

profile the thickness of both first-year and multi-year sea ice. It was from this<br />

study that the first assessment of the volume of oil which could pool up under sea<br />

ice was made. The radar profile depth data were calibrated against drill hole ice<br />

thickness measurements. An example of the radar sounding data as displayed on a<br />

graphic record is given in Figure 2. A more detailed description of the sounding<br />

system has been given in Kovacs [7].<br />

At Tigvariak Island a 20- by 150-m section of a recently plowed snow-free sea<br />

ice runway was profiled. Snow depths near the runway varied from 24 to 46 cm over a<br />

distance of 120 m. The mean depth was 34 cm with a standard deviation of 7 cm. Ice<br />

thickness was measured along 18 lines 1.1 m apart running parallel to the long axis<br />

of the runway. From the digitized radar ice thickness data cross sections were constructed<br />

(Fig. 3). The under-ice depressions seen in Figure 3 occurred at about 40-m<br />

intervals and in general could be correlated with the major snowdrift relief observed<br />

alongside the runway. The ice thickness data were also used to make a contour map of<br />

the under-ice relief (Fig. 4), in which ice thickness from 1.5 to 1.65 m is shown at<br />

contour intervals of 5 cm. The mean ice thickness was 1.55 m, with a standard deviation<br />

of 0.03 m. The shaded areas represent ice which is less than 1.55 m thick, i.e.<br />

where oil could be expected to accumulate. The black areas represent pockets where<br />

the ice is also less than 1.55 m thick but where surrounding thicker ice (white areas)<br />

might prevent the influx and accumulation of oil. Chances are, however, that some of<br />

these pockets would also fill with oil, as the surrounding ice appeared to be of no<br />

more than average thickness.<br />

For the 20- by 150-m runway area profiled, the quantity of oil which could be expected<br />

to pool up under ice of less than the mean thickness was found to be 0.0320<br />

m 3 /m 2 or 32,000 m 3 /km 2 •<br />

Analysis of the 18 ice profiles in 30-, 60-, 90-, 120- and 150-m-long segments<br />

resulted in the findings shown in Table I. This table shows that mean ice thickness<br />

data from an area 20 m wide and 30 m long are not representative of the total runway<br />

area, but that data from an area 60 m or more long do give a representative mean. An<br />

important finding is that a single traverse only 30, 60 or 90 m long is not reliable<br />

for determining the under-ice oil pooling capacity. For example, the storage volume<br />

above the mean thickness for each of the 18 30-m-long traverses varied from 18,300<br />

m3/km 2 to 56,300 m3/km 2• For the 6D-m-long segments the volume varied from 20,700 to<br />

46,800 m 3/km 2 and for the 90-m-long segments it varied from 25,000 to 39,800 m 3 /km 2 •<br />

Table I shows, as would be expected, that with increasing traverse length the standard<br />

deviation of the potential storage volume which can be determined from a single<br />

traverse decreases. The data in Table I are shown graphically in Figure 5, from which<br />

one can infer that a single ice thickness profile 150 m or more long would have provided<br />

data from which the potential under-ice oil pooling capacity of the local sea<br />

ice area could be determined with a high degree of confidence.<br />

914<br />

Table. I. Mean sea ice thickness and potential under-ice oil pooling<br />

volume for 18 ice profile segments obtained at Tigvariak Island.<br />

Profile length Mean thickness Std dev Mean vol Std dev<br />

(m) (m) (m) (m 3 /km 2 ) (m 3 /km 2 )<br />

30 1.497 0.026 33,200 7,660<br />

60 1.541 0.028 29,700 6,750<br />

90 1.539 0.036 31,100 3,540<br />

120 1.544 0.031 29,500 2,240<br />

150 1.546 0.031 32,000 1,590<br />

1.537* 31,100*<br />

* Mean of means.


Figure 7. Aerial view of undisturbed terrain at the Prudhoe Bay west dock before<br />

the area was cleared of snow and ice blocks. The rectangle encloses the<br />

area profiled for ice thickness. A refrozen crack about 2 rn wide runs between<br />

A and B.<br />

At the west dock site (Fig. 7) we graded off the snow and the uplifted ice<br />

blocks from the sea ice surface. The deformed ice typically was less than 1/2 m high<br />

and consisted of minor ice blocks 15 Cm thick randomly scattered in a sinuous line,<br />

i . e. the blocks did not form an ice pile per se. The profile area was 127 rn wide by<br />

160 m long . Seventy-eight 160-m-long profile lines 1.65 m apart were run. The radar<br />

data showed the mean ice thickness to be 1.83 m with a standard deviation of 0.15 m.<br />

The pocket volume above the mean thickness was 1237 m 3 • This translates into a potential<br />

oil pooling capacity of 60,500 m 3 /km 2 or nearly twice that of the ice area profiled<br />

at Tigvariak Island. This higher storage capacity resulted in part from a<br />

marked variation in snow cover depth, particularly around the uplifted ice, and in<br />

part from deformation features, including ridge keels and refrozen cracks, some about<br />

2 m wide, in which the ice was only about 1 m thick. Contour maps were constructed to<br />

provide insight into the under-ice relief . Figure 8 is a map showing where the ice<br />

thickness was less than (black and shaded areas) or more than (white areas) the mean .<br />

In the shaded areas oil would probably flow under the ice . The black areas are isolated<br />

by deeper surrounding ice (white areas) and might not be reached by the oil;<br />

they represent less than 5% of the total volume above the mean depth .<br />

When the relief map (Fig. 8) was placed over the near vertical aerial view of<br />

the undisturbed study area (Fig . 7) we found that the ice ridges in the northern portion<br />

of the study area corresponded to the white areas of the map, i . e . where thicker<br />

ice existed. In addition , where there were deep snowdrifts along the ridges the ice<br />

was thinner, as it was under some but not all of the snow sastrugi . The location of a<br />

refrozen crack about 2 m wide that passed thr ough the profile area from A to B was<br />

clearly evident on the contour map and indeed overlies this feature ' s surface manifestation,<br />

which is traceable in Figur e 7. One large snowdrift located at position C in<br />

Figure 8 was found to have thinner ice underneath.<br />

917


References<br />

1. Barnes, P.W., E. Reimnitz, L.J. Toimil and H.R. Hill (1979) Fast ice thickness<br />

and snow depth relationships related to oil entrapment potential, Prudhoe<br />

Bay, Alaska, 5th Int. Conf. on Port and Ocean Engineering under Arctic Conditions,<br />

Norwegian Institute of Technology.<br />

2. Cox, J.C., L.A. Schultz, R.P. Johnson and R.A. Shelsby (1980) The transport<br />

and behavior of oil spilled in and under sea ice, Arctec Inc., Report 460C.<br />

3. Greene, G.D., P.J. Leinonen and D. Mackay (1977) An exploratory study of the<br />

behaviour of crude oil spills under ice, the Canadian JOl1rnal of Chemical<br />

Engineering, Vol. 55.<br />

4. Hanson, A.M. (1980) The snow cover of sea ice during the Arctic Ice Dynamics<br />

Joint Experiment, 1975 to 1976, Arctic and Alpine Research, Vol. 12, No.2.<br />

5. Hibler, W.D. III (1980) Sea ice growth, drift, and decay, In: Dynamics of Snow<br />

and Ice Masses, ed. S. Colbeck, Academic Press Inc.<br />

6. Holtsmark, B.E. (1955) Insulating effect of a snow cover on the growth of young<br />

sea ice, Arctic, Vol. 81, No.1.<br />

7. Kovacs, A. (1977) Sea ice thickness profiling and under ice oil entrapment,<br />

9th Annual Offshore Technology Conference, Houston, Texas, Paper OTC2949.<br />

8. Kovacs, A. and R.M. Morey (1978) Radar anisotropy of sea ice due to preferred<br />

azimuthal orientation of the horizontal c-axes of ice crystals, Journal of<br />

Geophysical Research, Vol. 83, No. C12.<br />

9. Larson, L. (1980) Sediment-laden sea ice: Concepts, problems and approaches,<br />

Outer Continental Shelf Environmental Assessment Program, Special Bulletin<br />

No. 29, ed. D.M. Schell, Arctic Project Office, University of Alaska.<br />

10. Malcolm, J.D. (1979) Studies of oil spill behaviour under ice, Proc. Workshop<br />

on Oil, Ice and Gas, Toronto, Ontario.<br />

11. Matthews, J.B. (1980) - Under-ice current regimes in the nearshore Beaufort<br />

Sea, In: Beaufort Sea Winter Watch Ecological Processes in the Nearshore<br />

Environment, Outer Continental Shelf Environmental Assessment Program,<br />

Special Bulletin No. 29, ed. D.M. Schell, Arctic Project Office, University<br />

of Alaska.<br />

12. Reimnitz, E. and K. Dunton (1979) Diving observations on the soft ice layer<br />

under the fast ice at DS-ll in Stefansson Sound boulder patch, U.S. Geological<br />

Survey, Annual Report, 1979, Attachment D.<br />

13. Reimnitz, E. and R. Ross (1979) Lag deposits of boulders in Stefansson Sound,<br />

Beaufort Sea, Alaska, U.S. Geol. Survey Open File Report 79-1205.<br />

14. Roots, F. (1971) Shore fast sea ice, Canadian Polar Continental Shelf Project<br />

Report 71-2.<br />

15. Weeks, W.F. and A.J. Gow (1979) Crystal alignments in the fast ice of arctic<br />

Alaska, U.S. Army Cold Regions Research and Engineering Laboratory, CRREL<br />

Report 79-22.<br />

922


J. D. Malcolm<br />

Associate Professor<br />

A. B. Cammaert<br />

Manager, Engineering Studies<br />

ABSTRACT<br />

MOVEMENT OF OIL AND GAS SPILLS<br />

UNDER SEA ICE<br />

Faculty of Engineering<br />

Memorial University<br />

St. John's, Newfoundland<br />

Acres-Santa Fe Incorporated<br />

Calgary<br />

Canada<br />

Canada<br />

Laboratory studies related to the <strong>com</strong>plex behavior of oil-well blowout products<br />

under Arctic ice are described. The studies include the disposition of crude oil<br />

and gas bubbles under smooth ice and the effects of currents on crude oil behavior<br />

under undulating ice covers when the undulations contain gas.<br />

The potential for ice fracture by a well blowout is reviewed, along with a brief<br />

survey of literature on sea ice roughness. Hypothesis concerning the behavior of<br />

well blowout products under Arctic conditions are discussed.<br />

INTRODUCTION<br />

The potential for Arctic oil spills is increasing dramatically with the level of<br />

petroleum exploration and development acti vi ty. Response to an oil spill in the<br />

Arctic is expected to be slowed by the remoteness, accessibility, severe climate,<br />

and a lack of fundamental knowledge of oil spill behavior under Arctic conditions.<br />

Oil spilled in the Arctic is estimated by Boyd [1 J to remain there for a period of<br />

the order of 50 years because of the slow rate of biological degradation at near-<br />

zero temperatures. During its lifetime, the spill will be diffused widely by the<br />

highly dynamic pack ice. Speculation on the area effected by an Arctic oil spill<br />

has received continuous attention for almost 10 years, largely because of the poten­<br />

tially catastrophic consequences. A lowering of the natural albedo of the Arctic<br />

ice by a widely dispersed spill may lead to permanent removal of the Arctic Ocean<br />

923


ice pack [2] with dramatic reordering of climate and temperatures over the entire<br />

globe. Many other more localized consequences are also possible.<br />

The transport of oil under the ice, particularly from an uncontrolled subsea well<br />

blowout, would represent one of the most efficient mechanisms for diffusing oil over<br />

a large area. Almost all known spill cleanup techniques developed within the past<br />

decade will prove useless until the oil migrates to the top surface of the ice.<br />

Monitoring of the spills' progress is possible, if the spill forms a well-behaved<br />

coherent pool beneath the ice, but detection is unlikely if the spilled oil breaks<br />

up into many widely dispersed globules. It is the purpose of this paper to report<br />

on several laboratory studies related to the motion of well blowout products under<br />

Arctic ice. These studies have been conducted over the past 3 years. We have<br />

resisted the temptation to speculate on the implications of our laboratory results<br />

for the areal extent of an Arctic well blowout.<br />

THE POTENTIAL FOR ICE<br />

FRACTURE BY A WELL BLOWOUT<br />

An Arctic oil well blowout in ice-covered waters would release a m1xture of gas and<br />

oil which would rise to the bottom of the ice cover and spread out beneath it. The<br />

volume of gas released is expected to exceed the volume of oil by a factor of about<br />

150. As the gas moves out under the ice, driven by either existing currents in the<br />

area or the horizontal currents generated as the blowout plume impacts the ice<br />

cover, the gas will displace the water from beneath the ice cover. At some distance<br />

from the plume impact zone, the plume currents may be reduced to the point where the<br />

gas <strong>com</strong>es to rest, relative to the ice cover.<br />

The forces exerted by a trapped gas bubble under thin ice, which tend to lift and<br />

fracture the ice sheet, have been analyzed by Topham [3]. Assumptions in the model<br />

include isotropic elastic properties for the ice sheet, which in reality is noniso­<br />

tropic, exhibiting both plastic and elastic properties dependent on the past history<br />

and climatic conditions. The analysis considered the ice to be thin, and deflec­<br />

tions small, and fracture was shown to occur either at the bubble center or just<br />

beyond the bubble edge--depending on bubble depth, ice thickness, and material<br />

properties assumed for the ice.<br />

In general, the flattened spherical shape of a gas bubble is truncated at some angle<br />

of contact with the solid which supports the bubble, and the contact angle could be<br />

in the range from 0 degrees to 180 degrees, depending on the surface. In the case<br />

924


The shear zone or seasonal ice zone features the greatest degree of roughness since<br />

it is characterized by hummocks and ridges. Wadhams [7] reports on the topography<br />

of the Beaufort Sea ice cover giving mean ridge height and spacing from airborne<br />

laser profiles. Ridges can act to contain the area of spill, and occur with a mean<br />

spacing of 1.5/km to 2.1/km in the Beaufort Sea with heaving ridging up to 8/km.<br />

Keel depths can average 21.2 m for light ridging and slow drift, 26.8 m for heavier<br />

ridging and fast drift with a 10-year maximum depth of 32.8 m. Wadhams estimates a<br />

vertical roughness scale of approximately 5 em for ice floes between ridges.<br />

The ice cover features in the vicinity of a well blowout cannot be predicted.<br />

However, laboratory studies can provide insight into the physical processes involved<br />

and in particular, can test hypotheses about the relative importance of various<br />

factors. In what follows we describe small experimental spills under both smooth<br />

horizontal ice covers and undulating ice covers.<br />

THE DISPOSITION OF CRUDE OIL<br />

AND GAS BUBBLES UNDER ICE<br />

Laboratory studies consisting largely of photographic observations have been<br />

conducted on the disposition of finite quantities of oil and air under smooth hori-<br />

zontal ice, or rather, its physical analog--plate glass. Air bubbles, like oil<br />

drops, lack sufficient pressure to squeeze out the thin water film separating the<br />

bubble from the glass. The gas bubbles are thus guaranteed to possess a contact<br />

angle of 180 degrees and the bubble thickness under horizontal glass or ice is given<br />

by equation (1) or (2).<br />

An air bubble 40 to 50 mL was first formed under a glass plate submerged 10 to 15 rom<br />

below the free surface in a tank of water. Crude oil drops with volume 0.5 mL were<br />

released at 10- to 15-s intervals 25 em beneath the gas bubble.<br />

The arrival of the first drop of oil at the gas/water interface creates a miniature<br />

spill of gas and oil with volume ratio about 150:1. The oil drop was found to<br />

remain intact for several seconds, presumably waiting for the thin film of water<br />

surrounding it to drain away. The oil drop then coalesces with the gas/water inter­<br />

face and spreads over it to form either a very thin film extending to the edge of<br />

the bubble, or a thick lens of oil confined to the bottom of the gas bubble. In the<br />

case of spreading crudes, the film covering the bubble can be extremely thin,<br />

exhibiting some colors of the spectrum, but a thicker lens of crude near the origi­<br />

nal oil drop location remains.<br />

926


The arrival of a successive train of oil drops at the gas bubble interface produce a<br />

growing oil lens in the gas bubble after each drop coalesces with the lens. Drops<br />

that land near the edge of the gas bubble slide upward along the bubble contour<br />

since the bubble surface is not horizontal, but curved. As the drop slides along<br />

the bubble contour, the film of water preventing immediate coalescence is gradually<br />

draining away. Some drops slide off the gas bubble and onto the ice (or glass in<br />

our experiments) before coalescence takes place. In general, coalescence time is<br />

short for the first oil drop then increases when a thick lens is present in the<br />

gas/water interface. Coalescence time can range from 1 to 2 s to 15 to 20 s and<br />

longer. As the oil lens thickens to about 2 mm, the gas bubble thickness is corres­<br />

pondingly decreased. When a thick oil lens covers the bottom of the gas bubble, oil<br />

drops tend to slide off the bubble more readily, and in the process drag oil from<br />

the lens up the side of the gas bubble. Oil drops shed from the gas bubble accumu­<br />

late on the glass surface and can coalesce to form a puddle, which is connected to<br />

the lens on the bottom of the gas bubble. This occurs when the addition of oil,<br />

drop by drop has reduced the ratio of gas to oil close to 1:1.<br />

Figure 1 illustrates the appearance of a gas bubble under glass with a quantity of a<br />

nonspreading crude oil forming a lens in the bubble interface. The crude oil is a<br />

mixture of Trinidad (24.8 percent), BCF (56.2 percent) and Leona (19.0 percent)<br />

supplied by Ultramar Canada Ltd. Figure 2 shows another gas bubble with a spreading<br />

crude oil lens. The oil is Tijuana Light and was supplied by Imperial Oil Ltd.<br />

Figure 3 shows the lens formed by Guanipa crude, and a thin oil film surrounds the<br />

lens. In Figure 4, the lens size has been increased by additional Guanipa, and a<br />

number of oil drops have slid from the gas bubble contour to take-up positions<br />

beside it.<br />

The experiments just described confirm the two-dimensional analytical study of<br />

Topham [8J, which suggest several stable configurations of oil and gas beneath a<br />

planar ice sheet. OUr experiments provide additional insight into the physical<br />

mechanisms controlling the various stable configurations observed.<br />

The mobility of the gas bubbles beneath smooth glass or ice is such that great care<br />

in leveling the plate had to be exercised to ensure a stable bubble position. The<br />

bubble buoyancy exceeds that of crude oil by a order of magnitude, and earlier<br />

studies on the mobility of crude oil under ice [9J, demonstrate sustained oil motion<br />

along a surface tilted as little as 2 degrees to the horizontal plane.<br />

927


Under rough sea ice the gas bubbles are expected to find their way into undulations,<br />

displacing the water. Whether the undulations will be <strong>com</strong>pletely filled with gas<br />

depends on the gas volume available. The effect of currents on crude oil motion<br />

under undulating ice covers when the undulations contain gas is described below.<br />

EFFECT OF CURRENTS ON OIL AND<br />

GAS SPILLS UNDER SEA ICE<br />

Testing was carried out in the Acres' ice flume facility. The flume is a se1f-<br />

contained recirculating system measuring 12 m long, 1.2 m high and 1.2 m wide. The<br />

flume and recirculating line are fully insulated with 10 em of styrofoam. One wall<br />

incorporates a 5.5 m length of acrylic windows behind hinged insulation panels,<br />

allowing visual inspection of test events. Bottom windows were installed to permit<br />

viewing and photographic lighting. A vinyl liner was installed to prevent leakage<br />

and fouling of flume walls with oil.<br />

A variable-speed pump provides continuous flow from 0 to 0.3 m 3 /s. Velocities<br />

under the ice cover were measured with a miniature flow meter. The ice cover was<br />

formed by circulation of air at a minimum temperature of -20·C at a velocity of<br />

7.4 m/s over the water or ice surface.<br />

refrigeration system.<br />

Air was cooled by a conventional 21 kW<br />

The flume was filled with water to a depth of 0.6 m and a sufficient quantity of<br />

salt was added to the freshwater to bring the salinity to 32 pro mille. The water<br />

was cooled to the freezing point (approximately -1. S·C) over a period of 2 days.<br />

The miniature flowmeter was frozen in place in preparation for the test runs.<br />

Thickness measurement scales were frozen in the cover at half-meter intervals along<br />

the length of the flume to measure ice thickness at undulation crests and troughs.<br />

At the same time insulated 20-cm diameter gas/oil injection ports were frozen in the<br />

cover at the trough locations.<br />

Undulations were formed by placing 2.5-cm thick sheets of rigid polyurethane insula­<br />

tion on the crest areas of the undulations, halting the growth of the ice cover at<br />

the desired thickness. The uninsu1ated trough areas of the cover were allowed to<br />

continue growing until the desired undulation shape had been attained. Undulation<br />

wave lengths were 1 m long for the 15-cm deep undulations and 1.5 m long for the<br />

7.5-cm and 2-cm deep undulations.<br />

928


An air piston and curved delivery tube was used to release air through the injection<br />

ports directly below the center of the depressions in the ice cover. After the<br />

desired volume of gas (air was substituted for natural gas for safety purposes) was<br />

placed in the depression under the =ver, the required volume of oil was added<br />

(proportional to the volume of gas) by gravity feed through the curved delivery<br />

tube. The spreading and interaction of the oil and gas were observed under static<br />

conditions, and the oil was allowed to cool.<br />

The flow was gradually increased in steps. The flow was allowed to stabilize at<br />

each step and the behavior of the oil and gas was observed. Tests were performed<br />

for gas/oil ratios of 150:1 and 15:1. Undulation amplitudes tested were 15 em,<br />

7.5 em, and 2 em.<br />

Gas and oil were injected into the 15-cm undulation at a gas/oil ratio of 150: 1.<br />

Following the injection of 9 L of air, 60 mL of oil was injected. The undulation<br />

was approximately one-third full. When the oil was being injected it rose from the<br />

tip of the injection system at the bottom of the flume in a series of pinched off<br />

jets (pinched off in the form of pendent drops). The pendent drops were character­<br />

istically 1 em or 2 cm in diameter at the most, and they rose to the undersurface of<br />

the ice as a series of drops under the action of buoyant forces. The drops would<br />

impact in the ice surface and bounce and deform slightly, gradually roll up the<br />

underside of the curved ice surface to meet the air/water interface at the bottom of<br />

the air pocket. The amount of oil injected was not large (60 mL) and the oil pool<br />

collected at the upstream end of the air/water interface in the undulation. After<br />

the oil was injected the pump was turned on and the velocity increased gradually.<br />

As the flow increased to 14 em/s, the oil began to migrate downstream across the<br />

undulation. At 18 em/s, most of the oil had been herded against the downstream edge<br />

of the gas pocket under the ice. At 30 em/s, it was observed that the oil slick<br />

gradually circulated within the air/water interface, but the oil motion was (J"t<br />

particularly rapid; no oil drops were broken off from the oil slick and swept down­<br />

stream in the high-current flows and, in fact, no motion of the air within the<br />

pocket or waves on the air/water interface were observed whatsoever.<br />

Figure 5 illustrates the sequence of observations on the next test. Gas and oil<br />

were injected at a gas/oil ratio of 150: 1 into a second 15 cm undulation located<br />

upstream from the undulation used in the previous test. A larger volume of gas was<br />

injected in the depression (12 L of air with 80 mL of oil). Initially the air/water<br />

929


interface was at approximately two-thirds of the undulation depth. At a flow velo­<br />

city of 14 em/s, the oil slick started to migrate to the downstream edge of the gas<br />

pocket in the undulation. As the flow was increased in steps beyond 24 em/s, the<br />

oil was definitely herded against the downstream edge of the undulation. At<br />

27 cm/s, fine air bubbles were visible in the flow indicating entrainment of air<br />

from the flume headbox. At a flow of 30 em/s, the slick thickened against the<br />

downstream edge of the gas pocket against the ice. Air entrained in the flow began<br />

filling the undulation. With approximately 2 em of ice below the gas/water inter­<br />

face, the slick was forced out of the undulation. rippling of the oil/water inter­<br />

face and shedding of oil droplets from the slick were observed. The slick elongated<br />

and large droplets of oil were broken off. The oil moved slowly and erratically<br />

under the rough ice cover. The oil seemed to alternately stick to the rough ice<br />

surface and break free.<br />

Additional oil (540 mL) was added to the 15-cm undulation used previously to bring<br />

the gas/oil ratio to 15: 1 with the undulation half full. The oil slick <strong>com</strong>pletely<br />

covered the bottom of the gas pocket, but the slick thickness was less than 0.5 em.<br />

As the flow was increased to 30 em/s, no reaction to the flow was observed beyond<br />

migration of oil downstream and slight thickening of the oil slick. Undoubtedly,<br />

the oil slick was protected from the high flows by the depth of the undulation below<br />

the gas/water interface.<br />

A 7.5-cm undulation was filled two-thirds full with 30 L of air and 200 mL of addi­<br />

tional oil. The flow was increased in steps to 18 em/so at which point, significant<br />

herding of smaller slicks to form one large slick at the downstream edge of the<br />

depression occurred. The slick thickened to form a 7.5-mm thick lens. At 25 em/s,<br />

a steady supply of air from upstream entered the undulation and was vented to main­<br />

tain an air pocket thickness of 6 em, leaving a 1.5 em depth of ice (to the tip of<br />

the keel) below the air/water line. The slick rotated to align itself against the<br />

downstream edge of the air/water/ice contact line. At 30 em/s no significant change<br />

had occurred. At 44 em/s, l-mm diameter oil droplets were entrained in the flow.<br />

A 2-cm undulation was filled with 15 L of air and 100 mL of oil to make it half<br />

full. The slick configuration was as observed before; randomly distributed thin<br />

patches of oil.<br />

At a flow of 12 cm/s some herding of oil was observed at the downstream end of the<br />

shallow depression. Herding was more evident at 18 em/s, but the oil was still<br />

930


contained in the depression. At 21 cm/s, a droplet of oil was sheared away from the<br />

downstream edge of the main slick of oil which was starting to flow slowly under the<br />

ice out of the air filled part of the undulation. At 23 em/s, the shedding of drop­<br />

lets from the main slick was more pronounced. At 25 em/s the major portion of the<br />

patch had moved down out of the undulation and under the ice at a velocity of<br />

0.5 cm/s and was elongating and breaking up under the shearing action of the flow.<br />

At 30 cm/s, interfacial waves on the remaining oil slick were observed and shedding<br />

of 2-cm diameter droplets (larger than the previously observed droplets) was<br />

observed. The slick was 0.5 to 1 em thick. More air had collected in the undula­<br />

tion leaving a 0.5 em depth of ice below the air in the undulation.<br />

From the short series of tests performed it can be concluded that the spreading of<br />

oil under an ice cover, when released in the presence of gas, depends on the effec­<br />

tive configuration of the underside of the ice cover. The volume of gas and undula­<br />

tion geometry <strong>com</strong>bine to determine the behavior of the oil slick on the gas/water<br />

interface.<br />

For the meter-long undulations tested, if the gas volume is small or the undulation<br />

depth is large enough that the gas/water interface is more than about 5 em above the<br />

bottom of the undulation of lower ice surface, the oil slick seemed to be fully<br />

protected from the flow. The only effect on the oil slick at velocities as high as<br />

44 em/s was herding or migration of the oil slick to the downstream limit of the<br />

gas/water interface below the gas pocket.<br />

However, when the ice extended only 1.5 em below the gas/water interface, the oil<br />

slick was forced out of the undulation at a flow velocity of 44 em/s and migrated<br />

under the ice cover. When the ice cover extended 0.5 em to 1.0 em below the<br />

gas/water interface, a flow velocity of 25 em/s was able to force the oil out of the<br />

undulation. (It should be noted that the size of the ice lip was difficult to<br />

observe under a fluctuating water level.)<br />

A relationship between flow velocity and depth of ice below the gas/water interface<br />

can not be determined from the limited amount of data available and it is suspected<br />

that the slope at the water line and wave length of the undulation is equally<br />

important. Further investigation should be directed toward the configuration of the<br />

contact line between the gas pocket, water level and ice cover to determine the<br />

effect of the depth and slope of the ice surface that the oil must descend to be<br />

forced from the undulation at a given flow velocity.<br />

931


CONCLUSIONS<br />

In the event of a well blowout under Arctic ice, there is evidence that the blowout<br />

products will be spread out under the ice rather than fracture it. Laboratory<br />

studies reported in this paper indicate that gas will fill the under-ice undulations<br />

near the blowout zone, probably creating a continuous gas/water interface for the<br />

oil to spread on in the form of thin lenses. At a greater distance from the blow­<br />

out, the gas bubble will be fragmented and oil and gas will be<strong>com</strong>e trapped in the<br />

ice cover undulations. In this region the oil lenses will be partially protected<br />

from currents by the ice keels below the gas/water line in each undulation.<br />

Future studies directed toward understanding the blowout generated currents, and the<br />

action of a continuously discharged oil and gas mixture under rough ice are<br />

re<strong>com</strong>mended. Only when detailed topographic mapping of the local under-ice surface,<br />

together with detailed knowledge of local currents, be<strong>com</strong>es available for a particu­<br />

lar well site, then forecasting of the areal extent and ultimate disposition of well<br />

blowout products may be<strong>com</strong>e viable.<br />

ACKNOWLEDGMENTS<br />

The experiments on oil and air under glass were conducted by students C. R. Dutton,<br />

C. D. Elliott, J. Hillman at Memorial University with funding provided by the<br />

Natural Sciences and Engineering Research Council of Canada, and Imperial Oil<br />

Limited. The authors are also grateful to Canadian Marine Drilling Ltd., and the<br />

Environmental Protection Service for funding the flume study at Acres laboratories,<br />

and permitting timely release of the results.<br />

932


934<br />

FIGURE I. AIR BUBBLE UNDER GLASS WITH<br />

NON -SPREADING CRUDE OIL LENS AND<br />

DROPLETS, VIEWED AT OBLIQUE ANGLE .<br />

FIGURE 2. AIR BUBBLE UNDER GLASS WITH<br />

SPREADING CRUDE OIL LENS


FIGURE 3 . AIR BUBBLE UNDER GLASS WITH GUANIPA<br />

CRUDE OIL LENS AND DROPLETS, VIEWED<br />

FROM BELOW AT OBLIQUE ANGLE<br />

FIGURE 4. SAME AIR BUBBLE AS IN FIGURE 3 . OIL<br />

LENS IS LARGER. OIL DROPS BESIDE<br />

BUBBLE SLID BENEATH LENS<br />

935


936<br />

SALINE<br />

WATER<br />

INSULATED COVER<br />

INSU LATED BOTTOM<br />

0.45m<br />

0) FLUME CROSS-SECTION AND GAS POCKET<br />

b) OIL INJECTION<br />

WATER FLOW<br />

( 30 em I s )<br />

o<br />

d) OIL DROPLET SHEDDING<br />

•<br />

FIGURE 5 . MECHAN ISMS OF OIL SLICK<br />

MOTION IN FLUME TESTS<br />

0 . 6m<br />

UlllIllIllIDUfLLW1lJJJJJ I IIIIill<br />

WATER FLOW<br />

( UP TO 24 em / s)<br />

c) OIL HERDING<br />

WATER FLOW<br />

( > 30 em Is )<br />

e) OIL MIGRATION


NEED FOR REAL WORLD ASSESSMENT OF THE<br />

ENVIRONMENTAL EFFECTS OF OIL SPILLS<br />

IN ICE-INFESTED MARINE ENVIRONMENTS<br />

Gordon A. Robilliard, Senior Project Scientist Woodward-Clyde Consultants U.S.A.<br />

Michael Busdosh, Senior Staff Scientist Woodward-Clyde Consultants U.S.A.<br />

ABSTRACT<br />

The increase in oil-related activities in the arctic will result in one or more<br />

significant oil spills in ice-infested marine waters. These spills will pro-<br />

bably be persistent and difficult to contain or clean up. Most of the effort to<br />

date has dealt with the fate of oil and developing counter-measures. Relatively<br />

little has been done to assess the real-world environmental, especially ecological,<br />

impacts of the few spills that have occurred to provide a basis for predicting<br />

realistic impacts of future spills. The paucity of impact assessment data appears<br />

to be related to the logistic and safety aspects of sampling spills in ice-infested<br />

waters. The sparse information that is available suggests that the ecological<br />

consequences are not different (or more or less) significant than those resulting<br />

from temperate or tropical spills; in fact, the data from the latter spills are<br />

directly useful in predicting the types of impacts that may occur in polar areas.<br />

The rate and time scale of oil degradation and impacts seem to be longer in<br />

the arctic, primarily due to the cold. However, for at least some macro­<br />

invertebrates, biological processes such as growth, sexual maturation and<br />

others seem to take longer so the overall consequences are probably similar<br />

in all marine environments. With the increase in the oil exploration activities<br />

in arctic regions, ac<strong>com</strong>panied by the concerns of regulators and citizens over<br />

oil spills, increased effort must be placed on documenting the "real world" impact<br />

of oil spills in arctic waters. These data will provide a basis for predicting<br />

with confidence the realistic environmental effects of oil spills in ice-infested<br />

marine environments.<br />

937


INTRODUCTION<br />

The increase of oil exploration production and transportation in arctic<br />

and sub-arctic marine environments will result in one of more significant oil<br />

spills. The oil spilled in these cold, ice-infested or ice-covered waters is<br />

likely to be persistent, be<strong>com</strong>e widespread and be especially hard to contain or<br />

clean up [1). Most research and technology development efforts have been<br />

directed to identifying the chemical and physical fate of spilled oil and to a<br />

means of locating, containing and cleaning up spilled oil, especially when ice<br />

is present (2). While this effort is a necessary step in developing effective,<br />

efficient counter-measures, it is also important to decide if these counter­<br />

measures are required for environmental reasons and, if so, where, when and how?<br />

To answer these latter questions we need a better understanding of the actual<br />

environmental effects and consequences of oil spills in ice-infested marine waters.<br />

At present, it is nearly impossible on the basis of existing data from actual<br />

spills to make a realistic prediction about the ecological consequences of an<br />

oil spill in arctic marine environments. Assessments of such spills have<br />

generally been based on "worst case" scenarios while the "most likely case"<br />

scenario receives little attention. For example, the "standard" worst case<br />

scenario is a large spill of crude oil spreading over a large area at freeze-up<br />

in a biologically productive and sensitive habitat. Another worst case scenario<br />

involves a large spill occurring in the Lancaster Sound-northern Baffin Bay<br />

area during the summer when the murres are all flightless and would be unable<br />

to escape oiling. Furthermore, the environmental consequences of these worst<br />

cases in the marine environment have generally been based on impacts<br />

associated with spills in temperate, ice-free environments or, at best, with<br />

spills in northern environments where ice is present on the water for a few<br />

weeks at the most [3,4). Whether or not this is justified is simply not known<br />

with confidence. Very few actual spills in ice-affected environments have been<br />

publicly documented and thus there are few data to use as a basis for predicting<br />

impacts of hypothetical spills.<br />

In most hypothetical impact assessments of polar oil spills, the implicit<br />

assumption is that since oil is not rapidly weathered or biodegraded in arctic<br />

environments, it will have significantly greater ecological consequences than<br />

in warmer waters where oil weathers and biodegrades faster. That is, the<br />

ecological functioning of arctic environments is different somehow than it is<br />

in warmer waters. However, we submit that though oil may degrade slowly (i.e. over<br />

several years or growing seasons) in the arctic, many biological processes such<br />

as growth, sexual maturation, etc. involving individual marine organisms as<br />

well as populations also seem to operate slowly (i.e. on the order of decades).<br />

938


We hypothesize that the environmental consequences of oil spills in arctic<br />

marine environments may be no more (or less) significant ecologically than in<br />

similar habitats in warmer waters once differences in the rate and time scale<br />

for ecological processes are recognized.<br />

REVIEW OF SOME SPILLS IN ICE-INFESTED WATERS<br />

There have been a number of planned as well as accidental oil spills in<br />

ice-infested waters, polar and otherwise. However the information on size and<br />

cause of the spill or type and source of oil is often sketchy and for many<br />

spills, is available only through word of mouth making the accuracy of the<br />

information questionable. It seems clear though that most of the spills have<br />

been (l)accidental, (2) of small volume, and (3) from ships or machinery<br />

exploring for oil or engaged in marine transportation.<br />

In most cases, the major environmental assessment effort of accidental<br />

(and planned) spills has emphasized the chemical and physical fate of the oil<br />

and the efficiency of any countermeasures [1,2,3,5,6,7,8]. Carstens and<br />

Stendstad [8] found that the shore fauna in the immediate vicinity of a<br />

diesel spill was wiped out but that the birds and benthic fauna were apparently<br />

unharmed. However, the benthic fauna had moderately high levels of aromatic­<br />

hydrocarbons in the tissue 67 days after the spill was discovered. A gasoline<br />

and diesel fuel spill on the sea ice at Deception Bay resulted in a loss of<br />

about 50% of the nearshore bivalves and 5% of the pOlychaetes [5]. Fucus and mussels<br />

were more affected by the cleanup measure (burning) than by the oil. The overall<br />

short-term biological damage appeared slight and localized; no estimate of long<br />

term damage or recovery was made. Petersen [9] found that Bunker C spilled in<br />

Melville Bay off Greenland underwent little microbial degradation after 4 weeks<br />

and that weathering was slow due to low temperatures and low light levels. He<br />

reported no adverse biological impacts though he expected the oil might remain<br />

in the bottom sediments for some time.<br />

A few investigators have tested the effects of oil on arctic marine<br />

organisms in field experiements rather than in laboratory situations. Busdosh, Atlas<br />

and their colleagues [12,13,14,15] found that in general, amphipods, pOlychaetes<br />

and other benthic organisms not unexpectedly choose oil-free sediments over oiled<br />

ones and survival was better in the clean sediments. However, after several<br />

weeks to months the differences began to disappear and the organisms' selectivity<br />

was less definite. Busdosh et al also found that acute toxicity disappeared<br />

939


elatively rapidly (i . e. within a period of days to weeks) and that toxicity<br />

appeared to be associated with the lighter hydrocarbon fractions. In these<br />

cold waters bacteria tended to degrade the oil more slowly than in warmer<br />

waters, but they were nevertheless fairly efficient over longer periods .<br />

These results from field experiments are generally supported by numerous<br />

studies conducted in the laboratory under more controlled, though less<br />

environmentally realistic, conditions. We have not attempted to review that<br />

literature here.<br />

OIL SPILLS IN ANTARCTICA<br />

The antarctic benthic <strong>com</strong>munity is generally pristine relative to hydro­<br />

carbon pollution due to man's activities. However, in at least two areas where<br />

man has been active, hydrocarbon pollution has been found. Clark and Law [10]<br />

<strong>com</strong>pared aromatic and aliphatic hydrocarbon loads in benthic organisms<br />

(predators) from Signy Island, a pristine area, and King Edward Cove, South<br />

Georgia, a whaling camp abandoned in 1965 after 61 years of use. They found<br />

significantly higher levels of petroleum hydrocarbons in the benthic organisms<br />

at King Edward Cove, probably due to spills from the whaling ships in years gone<br />

by. Platt and Mackie [11] measured the high levels of aromatic hydrocarbons in<br />

sediments of King Edward Cove, but because they did not analyze other pristine<br />

antarctic sediments, they attributed the high levels to worldwide distribution of<br />

pollutants rather than to local pollution.<br />

At Winter Quarters Bay, McMurdo Sound, Antarctica in August-September 1974,<br />

one of us (GAR) observed and photographed (Figure 1) a large bed (about 75m by 25m)<br />

940<br />

Figure 1. Bivalve ( Laternula ) shells in Winter Quarters<br />

Bay, McMurdo Sound, Antarctica


of clam shells, piled several deep on the surface of the bottom. The area is<br />

a shallow depression about ISO-200m long by about 100m wide, at about the 2Sm<br />

depth, and surrounded by a rim about 3-l0m high. The shells, probably Laternula,<br />

were lO-20cm long, and both valves were often side by side. It appeared as though<br />

the clams had been dug up, died, and the shells left in place. However, few<br />

shells were broken indicating that they were not physically removed such as can<br />

occur when icebergs gouge the bottom . We also noted that there was a<br />

considerable amount of trash on the bottom, primarily from the McMurdo Station<br />

garbage dump on the sea ice above and that the fauna was extremely depauparate<br />

<strong>com</strong>pared to physically similar areas nearby (Figure 2).<br />

Figure 2. Sponges, Anemones and other Benthic Organisms at<br />

Hut Point, Ross Island, Antarctica. Photo at same depth about<br />

1 km north from Figure 1.<br />

The surface sediment was reddish-brown to light brown silt, but the<br />

subsurface sediments were black and smelled strongly of petroleum hydrocarbons.<br />

Subsequently, (December 3-6, 1974) samples were obtained from this area as well<br />

as two other sites, one in 3m of water nearshore in Winter Quarters Bay and the other<br />

across McMurdo Sound at New Harbour (30m depth). There was no obvious smell of oil<br />

in the sediments at either place. These three samples were analyzed for<br />

hydrocarbons to determine the amount and type of hydrocarbon fractions present.<br />

The sediments from the clam bed in Winter Quarters Bay contained approximately<br />

0.23% petroleum hydrocarbon by dry weight of sediment. Although a small amount<br />

of the hydrocarbons apparently was biogenic in origin, most of it appeared to be<br />

lubricating oil and possibly heavy residual or Bunker C fuel. No diesel fuel was<br />

present. The hydrocarbon content of the other two samples was approximately<br />

941


0.02 and 0.03%, respectively, for New Harbour and Winter Quarters Bay stations<br />

and most of it was biogenic in origin.<br />

There seems to be little doubt that the presence of a large number of dead<br />

clams and the high level of petroluem hydrocarbons in the sediments are related.<br />

We hypothesize that the petroleum hydrocarbons were introduced to the environment<br />

rather suddenly although we do not know the exact mechanism. It is possible<br />

that they were in containers that were dumped overboard as part of the trash<br />

that is piled in the McMurdo garbage heap, much of which gets dumped into Winter<br />

Quarters Bay when the ice breaks up. It may have also been discharged overboard<br />

during the offloading of fuel from the tanker vessels to the shore-based storage<br />

tanks each summer; these tankers anchor directly over the clam bed area. After<br />

the oil was introduced into the substratum, it probably reached levels high<br />

enough to affect the clams and they attempted to escape the area by crawling<br />

out of and on top of the sediment. They were unable to escape from the area and<br />

finally died. This would explain why most of the shells were unbroken. It is<br />

difficult to tell how long ago the clams died, but several years (1978) later divers<br />

reported that there was no apparent change in the <strong>com</strong>munity and the sediments<br />

still smelled strongly of oil. We expect that it will be several years at<br />

least before clams and other deep burrowing forms will once again inhabit this<br />

area, though polychaetes and amphipods are abundant in the thin uncontaminated<br />

surface sediment layer (Oliver, personal <strong>com</strong>munication).<br />

CONCLUSIONS<br />

The general conclusions from these and several ongoing studies including<br />

the BIOS program at Pond Inlet, Baffin Island, under the auspices of the Arctic<br />

Marine Oilspill Programme, seem to be three. First, quantitatively sampling and<br />

assessing the biological consequences of oil spills in ice-infested marine en­<br />

vironments is extremely difficult, largely due to the logistic and safety aspects<br />

of getting to and then working in the ice-infested environment. Second, the<br />

impacts do not appear different, or more or less significant, than those in more<br />

temperate areas; that is, the information based on many, well-documented spills in<br />

sub-arctic, temperate and tropical waters seems to be applicable in principle<br />

though not always in detail. Third, the physical, chemical and ecological<br />

degradation of oil in these cold environments is slower than in temperate areas<br />

so that the time periods over which toxic effects may occur and <strong>com</strong>munity recover<br />

is delayed are longer. However, this may not be ecologically significant in the<br />

sense that normal ecological processes often occur more slowly and the final<br />

<strong>com</strong>munity response may be similar albeit slower than that described in ice-free<br />

942


environments. Only better documentation of actual spills will provide the final<br />

test for this hypothesis and a basis for predicting with some confidence the<br />

realistic impacts of and recovery from oil spills in ice-infested marine waters.<br />

ACKNOWLEDGMENTS<br />

We are grateful to several colleagues, especially Jon Percy and Eric<br />

Schrier, who provided information about the few arctic oil spills and to<br />

John Oliver who first brought our attention to the presence of oil in the<br />

Winter Quarters Bay sediments and provided the sediment samples for analysis. We<br />

are also grateful to Ed Owens and Jim Sartor for reviewing the paper. We<br />

acknowledge Woodward-Clyde Consultants for providing the typing, graphic and<br />

financial support to prepare this paper.<br />

REFERENCES<br />

[1]. Anon. 1980. An Oilspill in Pack Ice. C-Core Contract Report<br />

(Contract No. 055 79-00007) to Environment Canada, Environmental<br />

Protection Service, Ottawa.<br />

[2]. D. Mackay & S. Paterson (eds.). 1979. Oil, Ice and Gas. Publication<br />

No. EE-14, Institute of Environmental Studies, University of Toronto,<br />

Toronto.<br />

[3]. Arctec, Inc. 1977. Bouchard #65 Oil Spill in Ice-Covered Waters of<br />

Buzzards Bay. Prepared for Alaskan OCSEAP, National Oceanic and<br />

Atmospheric Administration, Boulder, Colorado.<br />

[4]. J. Vandermeulen (ed.) 1978. <strong>Proceedings</strong> of SympOSium on Recovery<br />

Potential of Oiled Marine Northern Environments. Journal Fish, Research<br />

Board of Canada. 35 (5).<br />

[5]. R. O. Ramseier, G. S. Gantcheff, and L. COlby. 1973. Oil Spill at<br />

Deception Bay, Hudson Strait. Scientific Series No. 29, Inland Waters<br />

Directorate, Environmen't Canada.<br />

[6]. F. G. Barber. 1970. Oil Spills in Ice: Some Cleanup Options. Arctic<br />

23:285-286.<br />

[7]. F. G. Barber. 1971. An Oiled Arctic Shore. Arctic 24:229.<br />

[8]. T. Carstands and E. Sendstad. 1979. Oil Spill on the Shore of an<br />

Ice-Covered Fjord in Spitsbergen. POAC 79, Proceeding of Port and<br />

Ocean Engineering Under Arctic Conditions Conference.<br />

943


[9]. H. K. Petersen. 1978. Fate and Effect of Bunker C Oil Spilled by the<br />

[10] .<br />

[ll] .<br />

[12] .<br />

[13] .<br />

[14] .<br />

[15] .<br />

944<br />

USNS Potomac in Melville Bay-Greenland 1977. Proc. Conference on<br />

Assessment of Ecological Impacts of Oil Spills. AIBS. Keystone, Colorado.<br />

A. Clarke and R. Law. 1981. Aliphatic and Aromatic Hydrocarbons in<br />

Benthic Invertebrates from Two Sites in Antarctica. Marine Pollution<br />

Bulletin. 12(1):10-14.<br />

H. M. Platt and P. R. Mackie. 1979. Analysis of Aliphatic and Aromatic<br />

Hydrocarbons in Antarctic Marine Sediments Layers. Nature (London)<br />

280:576-578.<br />

R. M. Atlas and M. Busdosh. 1976. Microbial Degradation of Petroleum<br />

in the Arctic. <strong>Proceedings</strong> of the Third International Biodegradation<br />

Symposium. Applied Science Publishers Ltd. London pp. 79-85.<br />

R. M. Atlas, A. Horowitz and M. Busdosh. 1978. Prudhoe Bay Crude Oil<br />

in Arctic Marine Ice, Water and Sediment Ecosystems: Degradation and<br />

Interaction with Microbial and Benthic Communities. Journal Fish,<br />

Research Board Canada. 35(5):585-590.<br />

M. Busdosh, K. W. Dobra, A. Horowitz. S. E. Neff and R. M. Atlas.<br />

1978. Potential Long-term Effects of Prudhoe Bay Crude Oil in Arctic<br />

Sediments on Indigenous Benthic Invertebrate Communities. Proc.<br />

Conference on Assessment of Ecological Impacts of Oil Spills. AIBS Keystone,<br />

Colorado.<br />

M. Busdosh. 1978. The Effects of Prudhoe Crude Oil Fractions on the<br />

Arctic Amphipods Boeckosimus affinis and Gammarus zaddachi.<br />

Dissertation submitted to Department of Biology, University of<br />

Louisville, Louisville, Kentucky III pp.


istics of surface water temperature distribution These are: 1) large scale warm water<br />

patches are formed in the coastal region with the width extending beyond 1.5 km from<br />

the shoreline, and 2) the temperature distribution pattern is classified into three<br />

categories; that is, the first and second ones are closely correlated to the northward<br />

and southward currents and the third one is the transitional stage from the northward<br />

to the southward or vi ce versa. In Fi gure 7, Case A 1 demonstrates a typi ca 1 temperature<br />

pattern induced by the northward current, Cases A2 and A3 represent transitional stages<br />

and Case A4 demonstrates a typical one induced by the southward current. That is to<br />

say, the observation period in September was fortunately set to catch the three stages<br />

of temperature distribution patterns. On the other hand, the pattern observed in<br />

December was only southward.<br />

Through the present investigations, it is realized that the heated water discharged<br />

through the outlets is convected by the large scale alternating current appeared in the<br />

coastal region and is diffused during the convective movement.<br />

The time required to finish the measurement of surface temperature distribution<br />

by using a boat was about 2 hours for one run. Hence a question was arised on the<br />

reliability of measuring technique. In order to check this problem, the airborne infrared<br />

scanning image was taken during the period of Case A4 in Figure 7. The both patterns<br />

derived are basically the same in spite of using different techniques.<br />

(3) Vertical distribution of water temperature: Figure 9 shows the vertical distribution<br />

of water temperature along the three lines set from shore to offshore at the south<br />

outlet, the Ottozawa River, and the Kumakawa River. The present measurement was made<br />

in the morning of September 16, when the southward current was predominant. Therefore<br />

the temperature condition was the same as in Figure 7 (d). From these diagrams it is<br />

recognized that the discharged warm water diffuses within a surface layer of 2 to 3 m<br />

thickness with a layer of temperature discontinuity. The surface temperature decreases<br />

with the distance from the outlet. Figure 10 indicates temperature records of a thermister<br />

chain set at the site of B-N in December 1979. Along the chain eleven thermisters<br />

were installed and the three records of thermisters at the depths of O.Sm, 2m<br />

and 8m from the surface were analyzed. Abrupt temperature rises at O.Sm and 2m<br />

depths are clearly shown in this diagram and are certainly caused by the warm water<br />

patch passing through the measuring site.<br />

(4) Mixing between the warm water and the circumference water: Another interesting<br />

result obtained in due course of measurements is the horizontal temperature discontinuity.<br />

Figure 11 shows the records of two thermisters set at O.Sm and 2.0m depths<br />

below the water surface on the boat and pulled along the measuring line parallel to the<br />

coastline at the distance of 1.5 km on December 14, 1979. These records indicate that<br />

949


the mixing of warm water with surrounding water is not active as we expected and a warm<br />

water front is used to be formed. The temperature difference reaches to 2 to 3 degrees<br />

C and the above front appears to be slick along which dust and foam are gathering.<br />

CONCLUSIons<br />

In the first part of this paper. the authors reviewed the general trend of energy<br />

consumption in Japan and pointed out that the tremendous amount of heated water has<br />

been discharged to the nearshore area and has given some environmental impacts to the<br />

coastal region. However the direct impact of heated water to the marine growth has not<br />

yet be clarified.<br />

On the other hand. great efforts have been devoted to carry out i ntensi ve fi eld<br />

observations. through which a number of information on the diffusion and dispersion of<br />

heated water discharged from the outlet of power stations. The authors also have done<br />

a series of field investigations in September and December 1979. at the Fukushima<br />

Nuclear Power Station, and have realized that the warm water is convected by the long<br />

oscillatory current with the predominant period of 2 to 3 days and diffused during the<br />

movement of warm water patch. The stated oscillatory current seems to be generated by<br />

the migrating front in spring and autumn. and by other unknown reasons.<br />

In order to understand the details of diffusion and disperson processes of warm<br />

water more precisely and to evaluate the environmental impact of warm water on marine<br />

life. synthetic investigations are re<strong>com</strong>mended to ac<strong>com</strong>plish in the near future. However<br />

some approaches have recently been tested to find out its possible influence. For<br />

example. research engineers at the Fukushima Fishery Experiment Station are trying to<br />

catch the behaviour of salmons who are <strong>com</strong>ing back to and going up their mother river<br />

(the Kuma River). the mouth of which is presently affected by the warm water discharged<br />

from the power station. Such joint efforts between engineering scientists and biologist<br />

are cordially appreciated to expand our knowledges on the present <strong>com</strong>plicated phenomena<br />

and to solve the practical problems.<br />

ACKtIOWLEDGEt1EIITS<br />

The present investigation was carried out under the Scientific Research Grant of<br />

the Ministry of Education. Science and Culture. Japan. The authors acknowledge to the<br />

personnel at the Tokyo Electric Company. who kindly permits them to use their valuable<br />

data.<br />

REFERENCES<br />

Nakamura, Y. (1979): Long period oscillation of coastal current. Chapter II. Section<br />

950


L'ACTION DES GLACES SUR LES LITTORAUX<br />

Jean-Claude DIONNE<br />

Departement de Geographie, Universite Laval, Quebec<br />

ABSTRACT<br />

The termglaaiel refers to all processes, forms, sediments or features related<br />

to drift ice action, including icebergs, in the various sedimentary environments.<br />

Glaciel processes occur on about 200 000 km of the world shoreline in both hemi­<br />

spheres. In Canada, approximately 90% of the marine and lacustrine shorelines are<br />

subjected to drift ice action. The subarctic regions and the mid-latitude temperate<br />

regions with cold winters, especially the macrotidal environments, are the most<br />

exposed to ice processes.<br />

Drift ice activity is relatively <strong>com</strong>plex and varies in intensity with latitude<br />

and the changing environment. In general, drift ice is considered an important agent<br />

of erosion, transportation, sedimentation and protection. Drift ice erodes shores<br />

in unconsolidated deposits, and picks up large quantities of sediments which are<br />

dispersed and scattered over various distances according to the melting rate of<br />

floes. Every year, millions of tons of sediments are displaced and distributed by<br />

ice along cold region shorelines giving them a particular aspect. In addition, ice­<br />

push action <strong>com</strong>monly destroys beaches and alters unconsolidated shorelines. From<br />

a practical point of view, drift ice often acts as a negative process destroying<br />

shore defences, walls and dykes made of unconsolidated material. Drift ice also<br />

constitutes a serious obstacle to coastal navigation and causes problems for access<br />

to harbours. In addition, drift ice <strong>com</strong>monly fills navigation channels and port<br />

basins with sediments. Drift ice is a universal process that should be considered<br />

seriously in planning the development of cold region shorelines and the design of<br />

new harbours and facilities.<br />

955


La sedimentation glacielle est certainement aussi importante que l'erosion,<br />

puisqu'elle en est la contrepartie. Son caractere positif n'est pas toujours sou­<br />

haitable; c'est Ie cas des apports de sediments dans les chenaux de navigation,<br />

les voies d'approche des installations portuaires et les bassins de mouillage pres<br />

des quais.<br />

La sedimentation glacielle presente plusieurs facettes suivant les milieux.<br />

II existe des differences fondamentales par exemple, entre Ie bas et Ie haut estran<br />

(plage), entre les rivages a pente faible et a pente raide, entre les rivages ro­<br />

cheux et ceux en materiel meuble. Ces differences ont ete mises en evidence dans<br />

diverses publications [22]. Rappelons simplement que si les glaces deposent leur<br />

charge detritique au hasard de la fonte, elle en abandonnent la plus grande partie<br />

dans les zones littorales et prelittorales.<br />

Au Quebec par exemple, dans les zones intertidales du Saint-Laurent et de la<br />

baie de James, les glaces abandonnent chaque annee plusieurs millions de tonnes<br />

de debris [19,27]. L'apport de blocs erratiques constitue probablement la manifes­<br />

tation la plus evidente. Sur la rive sud du Saint-Laurent, les estrans sont fre­<br />

quemment couverts de blocs cristallins provenant du Bouclier canadien situe sur la<br />

rive nord, entre 10 et 35 km de distance. Or, ces blocs reposent directement sur<br />

des sediments marins fins post-glaciaires de plusieurs metres d'epaisseur. La meme<br />

situation prevaut sur la cote orientale de la baie de James ou des centaines de<br />

milliers de cailloux capitonnent les estrans argileux, vaseux ou sableux. Le con­<br />

traste frappant entre ces elements grossiers et les substrats meubles qui les por­<br />

tent met en evidence Ie role majeur des glaces dans la sedimentation littorale des<br />

regions froides.<br />

Les glaces transportent aussi de grandes quantites de sediments dont la taille<br />

est inferieure a celIe des blocs (25 cm). On trouve frequemment, a la surface des<br />

estrans et des marais littoraux, des semis ou des ilots de galets, gravier, sable,<br />

vase ou argile qui contrastent avec les substrats qui les portent. Dans plusieurs<br />

baies ou rentrants des rives du Saint-Laurent et de la baie de James par exemple,<br />

les apports glaciels de toutes sortes sont si abondants qu'ils jettent parfois dans<br />

l'ombre l'action des vagues et des courants. La meme situation prevaut dans plu­<br />

sieurs autres regions froides dans Ie monde.<br />

Le role sedimentologique des glaces flottantes n'est pas restreint aux seuls<br />

apports de sediments. Les glaces influencent parfois profondement la sedimentation<br />

963


deterres et exposes au froid; il s'en suit une augmentation considerable de la<br />

mortalite. Les vegetaux n'echappent pas a 1a destruction par 1es glaces. Les<br />

plantes halophi1es et 1es algues sont fauchees ou arrachees par 1es glaces; cer­<br />

tains peup1ements peuvent meme etre dangereusement decimes. De plus, arbres et<br />

arbustes a la limite des hautes mers souffrent aussi de l'erosion glacielle.<br />

Ces quelques exemples montrent 1a necessite d'etudier serieusement le glacie1<br />

en milieu littoral, en particulier sur 1es cotes ou i1 existe des amenagements<br />

pour 1a navigation ou pour l'exp1oitation des richesses nature11es. C'est pour­<br />

quoi 1es deve10ppements en ce sens dans la mer de Beaufort et 1a plate-forme du<br />

Labrador tiennent <strong>com</strong>pte de plus en plus des effets nefastes des glaces flottantes<br />

et des icebergs [12, 53].<br />

CONCLUSION<br />

En conclusion, rappe10ns que si l'action des glaces est relativement bien<br />

connue dans ses gran des 1ignes, i1 reste encore beaucoup a faire loca1ement pour<br />

preciser ou quantifier cette activite et trouver des solutions efficaces pour<br />

contrer 1es effets negatifs des glaces. Heuseusement, de plus en plus de specia-<br />

1istes prennent conscience de ces prob1emes et oeuvrent ales resoudre. Rappe10ns<br />

aussi que 1es regions po1aires ne sont pas forcement 1es plus serieusement menacees.<br />

Les littoraux des regions subarctiques et temperees a hiver froid, plus peuplees<br />

et plus 1argement amenagees, souffrent davantage des effets negatifs des glaces.<br />

967


12. __ _ ,1971: <strong>Proceedings</strong> of the Canadian seminar on icebergs;<br />

Halifax, Canada Dept. National Defence, 175 p.<br />

13. CAREY, S.W. et AHMAD, N., 1961: Glacial marine sedimentation; dans<br />

Geology of the Arctic, G.O. Raasch, ed., Toronto<br />

Univ. Press, vol. 2, p. 865-894.<br />

14. CARSOLA, A.S., 1954: Microrelief on the Arctic sea floor; Bull. Amer. Assoc.<br />

Petroleum Geol., vol.38, p. 1587-1601.<br />

15. C.I.M.M., 1969: Ice seminar; Conference held in Calgary, May 6-7 1968,<br />

Can. Inst. Mining & Metallurgy, Sp. publ. no 10,<br />

110 p.<br />

16. DAVIS, R.A., 1973: Coast ice formation and its effects on beach sedimentation;<br />

Shore and Beach, vol. 41, p. 3-9.<br />

17. DEACON, G., ed., 1971: Antarctic ice and water masses; Proc. Symposium,<br />

Tokyo-1970; Bruxelles, Sci. Committee on Antarctic<br />

Research, 113 p.<br />

18. DIONNE, J.C., 1968: Morphologie et sedimentologie glacielles, cote sud du<br />

Saint-Laurent; Zeitsch. Geomorphologie, Sp. Bd.,<br />

no 7, p. 56-84.<br />

19. ,1970: Aspects morpho-sedimentologiques du glaciel, en particulier<br />

des cotes du Saint-Laurent; Quebec, Env. Canada,<br />

Lab. Rech. Forestieres, Rapp. Infor., QFX-9, 324 p.<br />

20. ,1971: £rosion glacielle de la slikke, estuaire du Saint-Laurent;<br />

Rev. Geomorph. Dynamique, vol. 20, p. 5-21.<br />

21. ,1972: Caracteristiques des schorres des regions froides, en<br />

particulier de l'estuaire du Saint-Laurent;<br />

Zeitsch. Geomorph., Sp. Bd.no 13, p. 131-162.<br />

22. ,1974a: Bibliographie annotee sur les aspects geologiques du<br />

glaciel; Quebec, Env. Canada, Centre Rech. For.<br />

Laurentides, Rapp. Infor. LAU-X-9, 122 p.<br />

23. ,1974b: Polished and striated mud surfaces in the St.Lauwrence<br />

tidal flats, Quebec; Can. J. Earth Sci., vol. II,<br />

p. 860-866.<br />

969


24.<br />

25.<br />

26.<br />

27.<br />

______ , ed. 1976a: Le glaciel; <strong>Comptes</strong> <strong>rendus</strong> P,remier Colloque Inter.<br />

Action geologique des glaces flottantes; Rev. Geogr.<br />

Montreal, vol. 30, 236 p.<br />

,1976b: L'action glacielle dans les schorres du littoral oriental<br />

de la baie de James; Cah. Geogr. Quebec, vol. 20,<br />

no 50, p. 303-326.<br />

,1978: Le glaciel en Jamesie et en Hudsonie, Quebec surbarctique;<br />

Geogr. Phys. Quater., vol. 32, p. 3-70.<br />

,1980: Les glaces <strong>com</strong>me agent littoral sur la cote orientale<br />

de la baie de James; <strong>Comptes</strong> <strong>rendus</strong> Conf. canadienne<br />

sur Ie littoral (Burlington, Ont.), Ottawa, Conseil<br />

Nat. Recherches du Canada, (CARERS), p. 80-92.<br />

28. DROUIN, M. et CARTER, D., 1973: Dynamique des glaces Ie long des rives du<br />

Saint-Laurent; Montreal, min. Travaux publics Canada,<br />

Etude des Rives du Saint-Laurent, Rapp. 43 p.<br />

29. GOLD, L.W. et WILLIAMS, G.P., 1968: Ice pressures against structures;<br />

Ottawa, Nat. Res. Council, Tech. Memo no 92, 247 p.<br />

30. GREEN, H.G., 1970: Microrelief of an arctic beach; J. Sed. Petrology,<br />

vol. 40, p. 419-427.<br />

31. GRIPP, K., 1963: Winter-PhHnomene am Meeresstrand; Zeitsch. Geomorpho.,<br />

vol. 7, p. 326-331.<br />

32. HIND, H.Y., 1875: The ice phenomena and the tides of the Bay of Fundy;<br />

Can. Monthly Nat. Rev., vol. 8, p. 189-203.<br />

33. HUME, J.D. et SCHALK, M., 1964: The effect of ice-push on arctic beaches;<br />

Amer. J. Sci., vol. 262, p. 267-273.<br />

34. KARLSSON, T., ed., 1972: Sea ice; Proc. Inter. Conf. on Sea Ice, Reykjavik,<br />

10-13 May 1971; Reykjavik, Iceland Nat. Res. Council,<br />

309 p.<br />

35. KINDLE, E.M., 1924: Observations on ice-borne sediments by the Canadian<br />

and other Arctic expeditions; Amer. J. Sci., vol. 7<br />

(5e ser.), p. 251-286.<br />

36. KINGERY, W.D., ed., 1963: Ice and snow; proterties, processes, and applications;<br />

Cambridge (Mass.), M. I. T. Press, 684 p.<br />

37. KNIGHT, R.J. et DALRYMPLE, R.W., 1976: Winter conditions in a macrotidal<br />

environment, Cobequid Bay, Nova Scotia;<br />

970 Rev. Geogr. Montreal, vol. 30, p. 65-85.


38. LAKTIONOV, A.F., 1957: The effects of ice upon shipping routes, sea and<br />

river ports and the means to <strong>com</strong>bat it;<br />

Proc. 19th Inter. Navigation Congress (London)-1957,<br />

sect. I, p. 177-217.<br />

39. LINDSAY, D.G., 1976: Sea-ice atlas of Arctic Canada, 1961-1968;<br />

Ottawa, Dept. Energy, Mines & Res., 213 p.<br />

40. LISITSYN, A.P., 1962: Bottom sediments of the Antarctic; Amer. Geophys.<br />

Union, Antarctic Res. Geophys. Monographs no 7,<br />

p. 81-88.<br />

41. ,1972: Sedimentation in the World Ocean; Tulsa (Oklahoma),<br />

Soc. Econ. Paleontologists & Mineralogists, Sp. Publ.<br />

no 17, 218 p.<br />

42. LYELL, C., 1854: Principles of Geology; New York, Appleton, 834 p.<br />

43. MARKHAM, W.E., 1968: Growth, break-up and movement of ice in Canadian<br />

coastal waters; Can. Inst. Min. Metal., Sp. Publ.<br />

no 10, p. 31-35.<br />

44. MILLER, J.A., 1966: The suspended sediment system in the Bay of Fundy;<br />

Halifax, Dalhousie Univ., Dept. Oceanogr., these<br />

maitrise, non publ., 105 p.<br />

45. MOIGN, A., 1976: L'action des glaces flottantes sur Ie littoral et les fonds<br />

marins du Spitsberg central et nord-occidental;<br />

Rev. Geogr. Montreal, vol. 30, p. 51-64.<br />

46. NICHOLS, R.L., 1961: Characteristics of beaches formed in polar climates;<br />

Amer. J. Sci., vol. 259, p. 694-708.<br />

47. OWENS, E.H. et McCANN, S.B., 1970: The role of ice in the Arctic beach<br />

environment; Amer. J. Sci., vol. 268, p. 397-414.<br />

48. PELLETIER, B.R., 1969: Submarine physiography, bottom sediments, and models<br />

of sediment transport in Hudson Bay; dans Earth<br />

Science Symposium on Hudson Bay; P.J. Hood, ed. Ottawa,<br />

Geol. Suv. Canada, Paper 68-53, p. 100-135.<br />

49. PELLETIER, B.R. et SHEARER, J.M., 1972: Sea bottom scouring in the Beaufort<br />

Sea of the Arctic Ocean; Proc. 24th Inter. Geol.<br />

Congress (Montreal-1972), vol. 8, p. 256-261.<br />

50. PREST, W.H., 1901: On drift ice as an eroding and transporting agent;<br />

Trans. Nova Scotia Inst. Sci., vol. 10, p. 333-344.<br />

971


51. PRESTWICK, J., 1886: Geology; Vol. I, Chemical and Physical; Oxford,<br />

Clarendon Press, 477 p. (cf. p. 186-192).<br />

52. REED, A. et MOISAN, G., 1971: The Spartina tidal marshes of the St. Lawrence<br />

estuary and their importance to aquatic birds;<br />

Natura1iste canadien, vol. 98, p. 905-922.<br />

53. RElMNITZ, E. et BARNES, P.W., 1974: Sea ice as a geological agent on the<br />

Beaufort Sea shelf of Alaska; dans The coast and shelf<br />

of the Beaufort Sea, J.C. Reed et J.E. Sater, eds.,<br />

Arctic Inst. North America, p. 301-351.<br />

54. REINECK, H.E., 1956: Wattenmeer im Winter; Senckenbergiana 1ethaea, vol. 37,<br />

p. 129-146.<br />

55. REX, R.W., 1955: Microrelief produced by sea ice grounding in the Chukchi<br />

Sea, near Barrow, Alaska; Arctic, vol. 8, p. 177-186.<br />

56. SHAPIRO, A. et SIMPSON, L.S., 1953: The effects of broken ice fields on<br />

water waves; Trans. Amer. Geophys. Union, vol. 34,<br />

p. 36-42.<br />

57. TARR, R.S., 1897: The arctic sea ice as a geologic agent;<br />

Amer. J. Sci., vol. 3, (4e ser.), p. 223-229.<br />

58. TAYLOR, R.B. et McCANN, S.B., 1976: The effects of sea and nearshore ice<br />

on coastal processes in Canadian Arctic Archipelago;<br />

Rev. Geogr. Montreal, vol. 30, p. 123-132.<br />

59. TYRRELL, J.B., 1910: Ice on Canadian lakes; Trans. Can. Inst., vol. 9,<br />

pt. 1, p. 13-21.<br />

60. U.R.S.S., 1960a: Atlas l'dov Baltiyskogo morya i pri1egayuschikh rayonov chasti<br />

(Atlas of the Baltic Sea and adjacent areas);<br />

Leningrad, Hydrometeoro10gica1 Press, 84 p.<br />

61. ,1960b: Al'born aerofotosnimkov 1edovykh obrazovanii na Moriakh;<br />

(Album de photographies aeriennes des glaces f10ttantes);<br />

Leningrad, Gidrometeoro10gicheskoe izd-vo, 221 p.<br />

62. VANNEY, J.R. et DANGEARD, L., 1976: Les depots glacio-marins actuels et anciens;<br />

Rev. Geogr. Montreal, vol. 30, p. 9-50.<br />

63. WARNE, D.A., 1970: Glacial erosion, ice rafting, and glacial marine sediments;<br />

Amer. J. Sci., vol. 269, p. 276-294.<br />

972


64. WASHINGTON, 1958a: Arctic Sea ice; Proc. on Arctic sea ice, (Easton,<br />

Maryland, 24-27 fev., 1958), Washington, D.C.,<br />

Nat. Acad. Sci. & Nat. Res. Council, Publ. 598, 271 p.<br />

65. ,1958b: Oceanographic atlas of the Polar seas; Pt. I, Antarctic,<br />

70 p.; Pt. II, Arctic, 149 p.; Washington, D.C.,<br />

U.S. Hydrographic Office, Publ. 705.<br />

66. ,1968: Oceanographic Atlas of the North Atlantic Ocean;<br />

section III, Ice; Washington, D.C., U.S. Naval<br />

Oceanographic Office, Publ. no 700, 157 p.<br />

67. WASHBURN, A.L., 1979: Geocryology. A survey of periglacial processes and<br />

environments; London, Edward Arnold, 406 p.<br />

68. WEEKS, W. et ASSUR, A., 1967: The mechanical properties of sea ice;<br />

Hannover (New Hampshire), CRREL, Rapp. 2-C3, 80 p.<br />

69. W.M.O., 1970: Sea ice nomenclature (Terminology, codes and illustrated<br />

glossary); Geneva, World Meteorol. Organization,<br />

Publ. 259, 147 p.<br />

70. ZENKOVICH, V.P., 1967: Processes of coastal development;<br />

New York, Wiley-Interscience, 738 p.<br />

71. ZUMBERGE, J.H. et WILSON, J.T., 1953: Effect of ice on shore development;<br />

Proc. 4th Conf. Coastal Eng. (Berkeley, Calif.),<br />

p. 201-206.<br />

973


John R. Harper<br />

and<br />

E.H. Owens<br />

ANALYSIS OF ICE-QVERRIDE POTENTIAL<br />

ALONG THE BEAUFORT SEA COAST OF ALASKA<br />

WOODWARD-CLYDE CONSULTANTS<br />

16 Bastion Square<br />

Victoria, B.C.<br />

vaw 1H9<br />

ABSTRACT<br />

Ice override, the process of sea ice thrusting landward across beaches and bar­<br />

rier islands, has been identified as a potential hazard to drilling and exploration<br />

activities along the Beaufort and Chukchi Sea coasts of Alaska. In order to better<br />

define the risk associated with ice override as an envircnmenta1 hazard, the objec­<br />

tives of this study were to (1) estimate the frequency of override occurrence along<br />

the Beaufort and Chukchi Sea coasts, and (2) delineate areas or regions of high ice­<br />

override potential. The results are based on an objective analysis of vertical aerial<br />

photographs and indicate that ice override occurs infrequently along both the Beaufort<br />

and Chukchi Sea coasts of Alaska. Evidence of 11 override events was observed on the<br />

611 km of Beaufort Sea coasts that were surveyed. Evidence of 9 override events was<br />

noted on the 970 km of Chukchi Sea coasts that were surveyed. Return periods for ice­<br />

override events, obtained by <strong>com</strong>bining the frequency of occurrence estimates in a<br />

probability model, are estimated at 93 years for the Beaufort Sea coast and 215 years<br />

for the Chukchi Sea coast. Return periods for the more severe override events are es­<br />

timated at 341 years and 323 years for the Beaufort and Chukchi Sea coasts respect­<br />

ively. Within the framework of the overall probability of occurrence of ice override,<br />

it is recognized that some locations may have higher levels of override potential.<br />

The study delineates three shoreline segments, two on the Chukchi coast and one on the<br />

Beaufort coast immediately east of Barrow, as having a greater potential of override<br />

occurrence, and suggests that the coastline east of the Colville River has a slightly<br />

lower level of override potential.<br />

974


INTRODUCTION<br />

Ice override is the process by which sea ice moves landward across the shore<br />

zone as an unbroken sheet and penetrates substantial distances inland from the high­<br />

water line. As such, ice override represents a potential hazard to structures placed<br />

on or near the shore. It is important to distinguish iae override from iae push,<br />

which also involves the landward movement of ice but which is limited to the zone of<br />

normal wave activity (usually


along Tapkaluk Island (20 km east of Barrow), where l-m thick ice overrode approxi­<br />

mately 3 km of coast and penetrated inland as much as 150 m [2]. This override event<br />

occurred during mid-winter (January, 1978) and is thought to have resulted from a<br />

large shore lead closing against the shore. Other more severe events have been docu­<br />

mented in the Canadian arctic, where one massive override of l-m thick ice penetrated<br />

as much as 185 m inland from +he high-water line [8]. Even thin ice «20 cm) can<br />

penetrate substantial distances inland, up to 100 m; however, analysis of the re­<br />

ported events suggests that a 30-60 m override-penetration distance is more typical<br />

[3]. In plan form, the ice often may appear as long tongues or fingers, as the pene­<br />

tration distance often exceeds the alongshore width of the override.<br />

There is no particular site condition which is consistently associated with lo­<br />

cations of ice-override, although Owens and McCann [6] note that in the Canadian<br />

arctic, ice override tends to be concentrated near major promontories along the coast.<br />

Override events have occurred in areas of low nearshore gradients of moderate offshore<br />

ice regimes and where offshore submarine bars are present.<br />

The existing reports on ice-override events provide little information on either<br />

the frequency of ice override as a process, or the type of site conditions that fa­<br />

vovr ice override. It is difficult to establish the frequency from the existing in­<br />

formation, as there exists an element of uncertainty whether events have simply been<br />

unnoticed in the past, or if ice override is actually a rare event (both in time and<br />

space).<br />

AIR PHOTO ANALYSIS<br />

Air photo interpretation was used to develop an ice-override data base that<br />

could be used to define the frequency of events. Low-level air photos (Tables 1<br />

and 2) were examined for direct evidence (i.e., ice on the beaches and backshore) and<br />

indirect evidence of ice-override events (i.e., ice-thrust scars on the beach, sedi­<br />

ment push-piles or gravel mounds, striations in the beach sediments). Where evidence<br />

of an ice override was noted, relevant information on the override characteristics<br />

(penetration distance, alongshore width, etc.) and site characteristics (foreshore<br />

slope, distance to 20 m contour, offshore ice-regime severity) were noted (see<br />

Tables 3 and 4).<br />

The survey technique and the resulting data base incorporate limitations with<br />

regard to the interpretations that can be made. The frequency values may be an<br />

underestimate, as (1) some of the override events may not be visible on the photos,<br />

either because the remnant scars were too small to be resolved, or because the beaches<br />

976


The air-photo survey data were used to quantify the risk-evaluation analysis.<br />

The data were <strong>com</strong>bined into a probability model to estimate the likelihood of an over­<br />

ride event affecting a given section of shoreline in a given length of time. The re­<br />

sults show that the probability of override occurrence along a 0.5-km segment of shore<br />

in a 20-year period ranges between 0.056 and 0.192 for the Beaufort Sea coast, and<br />

0.060 and 0.088 for the Chukchi Sea coast. The lower value of each probability range<br />

represents a probability associated with the more severe events. The corresponding<br />

return periods for the occurrence of ice override range from 93 to 341 years for the<br />

Beaufort Sea coast, and 215 to 323 years for the Chukchi Sea coast. The longer re­<br />

turn period of each range is associated with the more severe events.<br />

REGIONAL VARIATIONS IN OVERRIDE<br />

It should be noted that the frequency estimates (and associated probabilities)<br />

are average estimates for a random segment of coastline. The actual estimates for<br />

specific segments of coastline may be higher or lower than the average estimates,<br />

depending upon temporal and spatial factors applicable to the given segments.<br />

Beaufort Sea Coast<br />

The distribution of override events shows a relatively high concentration in the<br />

coastal segment immediately east of Point Barrow (Fig. 1). Six (54 percent) of the<br />

observations were concentrated on 27 percent of the surveyed shoreline. Of special<br />

interest is that 5 of the 11 events were identified on Tapkaluk Island (Fig. 1, events<br />

6 and 7), Martin Island (events 1 and 8), and Igalik Island (event 9), which are the<br />

islands where three large 1978 overrides occurred [2].<br />

Chukchi Sea Coast<br />

The distribution of events shows a distinct concentration of events near Point<br />

Barrow, and also in the vicinity of Point Lay (Fig. 2); 78 percent of the events were<br />

concentrated within these two segments, which <strong>com</strong>prised 25 percent of surveyed shore­<br />

line (11 percent in the Point Lay area and 14 percent in the Point Barrow area). The<br />

concentration of events near Point Lay is likely associated with a locally more in­<br />

tense offshore-ice regime [7] and a slight flexure in the coastline.<br />

Discussion<br />

It is clear that considerable spatial variation exists within the overall fre­<br />

quency estimates of ice override on both the Beaufort and Chukchi Sea coasts. Both<br />

980


00<br />

.... '"<br />

..<br />

AUF 0 T 5 E<br />

Figure 1. The <strong>com</strong>bined distribution of published • and surveyed CD ice override events<br />

along the Beaufort Sea coast; event numbers are keyed to Table 3.


Figure 2. The <strong>com</strong>bined distribution of published .. and surveyed CD ice override events<br />

along the Chukchi Sea coast; event numbers are keyed to Table 4.


the published and surveyed override data identify three zones of high override poten­<br />

tial - one on the Beaufort Sea coast immediately east of Barrow, a second near Barrow<br />

on the Chukchi Sea coast, and a third near Point Lay on the Chukchi Sea coast. The<br />

trends are especially important in terms of an interpretation of previous override re­<br />

ports. The implication is that the Barrow region, both to the east and southwest of<br />

Barrow, is not representative of the Alaskan coast as a whole, and that observations<br />

on both the frequency and severity of override events should not be extended to other<br />

sEctions of the coast without qualifications.<br />

The ice-override data base is not sufficiently large to delineate site condi­<br />

tions which are conducive to ice override, although there appears to be a general as­<br />

sociation of high override frequency with zones of intensive offshore ice movements<br />

and with large-scale promontories not protected by offshore shoals. Intuitively, one<br />

would expect that the following factors would increase the probability of ice override<br />

in a particular area: severe offshore ice regimes, steep offshore and nearshore grad­<br />

ients, the absence of offshore bars and shoals, and low backshore elevations.<br />

SUMMARY AND CONCLUSIONS<br />

Ice override is known to occur throughout the arctic, and has been documented<br />

along the Alaskan arctic coast. Ice override is capable of penetrating substantial<br />

distances inland from the coast, up to 150 m in some cases, and as such must be re­<br />

garded as a potential environmental hazard to structures placed near the shore zone.<br />

The level of risk associated with this process along the Alaskan arctic coast is rela­<br />

tively low, because ice-override is rare both in time and space. It should be empha­<br />

sized, however, that these results are derived for the Alaskan arctic coast, and ice<br />

override is a significant modifying process in other areas of the arctic (e.g., the<br />

south coast of Viscount Melville Sound is <strong>com</strong>pletely dominated by ice-override scars).<br />

The following specific conclusions may be drawn from the study:<br />

(1) An objective inventory of ice-override events, derived from<br />

vertical air-photo interpretation, provides an estimated frequency<br />

of ice override as 0.018 events/km of shoreline for the<br />

Beaufort Sea coast, and as 0.009 events/km of shoreline for<br />

the Chukchi Sea coast. The <strong>com</strong>bined frequency for the entire<br />

Alaskan arctic coast is 0.013 events/km.<br />

(2) The probability of override in any given O.s-km section of<br />

coast is low. For the Beaufort Sea coast, the probability<br />

of override occurrence in a 20-year time period ranges from<br />

0.19 to 0.056, with corresponding return periods of 341 to 93<br />

years. For the Chukchi Sea coast, the probability of override<br />

during a similar 20-year interval ranges from 0.06 to<br />

0.088, with corresponding return periods of 323 to 215 years.<br />

(3) Analysis of the vertical aerial photographs has tentatively<br />

defined three zones of high override potential. The zones are<br />

%3


located (a) near Point Lay on the Chukchi Sea coast, (b) near and including<br />

Point Barrow on the Chukchi Sea coast, and (c) on the segment<br />

of shoreline immediately east of Point Barrow (to Cape Simpson) on the<br />

Beaufort Sea coast. The important implications are that the Barrow<br />

area is not typical of the Alaskan arctic coast as a whole, in terms of<br />

override potential, and that observations from the Barrow area on both<br />

the frequency and the severity of events should not be extended to<br />

other areas without qualification.<br />

ACKNOWLEDGEMENTS<br />

Funding for this study was provided through a contract to Woodward-Clyde Consul­<br />

tants from Chevron, U.S.A., who also gave permission to publish the results.<br />

Ram Kulkarni of Woodward-Clyde Consultants performed the probability analysis.<br />

REFERENCES<br />

[1] A1esta1o, J., and Haikio, J., 1976. Ice features and ice-thrust shore forms at<br />

Luodonse1ka, Gulf of Bothnia in winter 1972-1973. Fennia 144, Geographical<br />

Society of Fin1arld, Helsinki, 24 p.<br />

[2] Hanson, A., Metzner, R., and Shapiro, L., 1978. Ice shove in the Point Barrow<br />

area. Arctic Project Bulletin #22, (OCSEAP), p. 4-8.<br />

[3] Harper, J.R., 1980. Analysis of ice override potential along the Beaufort Sea<br />

coast of Alaska. Woodward-Clyde Consultants, Anchorage, Alaska, unpublished<br />

proprietary report to Chevron, U.S.A., San Francisco, 114 p.<br />

[4] Kovacs, A., and Sodhi, D.S., 1979. Ice pile-up and ride-up on arctic and subarctic<br />

beaches. Proc. of 5th International Conference on Port and Ocean Engineering<br />

under Arctic Conditions, Norwegian Institute of Technology,<br />

Trondheim, p. 127-146.<br />

[5] ,1979. Shore ice pile-up and ride-up, field observations, models,<br />

theoretical analyses. O.N.R. Workshop on Problems of the Seasonal Sea Ice<br />

Zone, Naval Postgraduate School, Monterey, California, 84 p.<br />

[6] Owens, E.H., and McCann, S.B., 1970. The role of ice in the Arctic beach environment<br />

with special reference to Cape Ricketts, southwest Devon Island,<br />

N.W.T., Canada. American Journal of Science, 268(5), p. 397-414.<br />

[7] Stringer, W.N., 1978. Morphology of Beaufort, Chukchi and Bering Seas nearshore<br />

ice conditions by means of satellite and aerial remote sensing. NOAA­<br />

OSCEAP, Boulder, Colo., Annual Report of Principal Investigators,<br />

Vol. I (218 p) and Vol. II (576 p).<br />

[8] Taylor, R.B., 1977. The occurrence of grounded ice ridges and shore ice piling<br />

along the northern coast of Somerset Island, N.W.T. Arctic, 31(2),<br />

p. 133-149.<br />

984


Austin Kovacs, Research Civil Engineer<br />

Devinder S. Sodhi, Research Hydraulic Engineer<br />

SEA ICE PILING AT FAIRWAY ROCK,<br />

BERING STRAIT, ALASKA:<br />

OBSERVATIONS AND THEORETICAL ANALYSES<br />

U.S. Army Cold Regions<br />

Research and Engineering<br />

Laboratory<br />

Abstract<br />

Information on sea ice conditions in the Bering Strait and the icefoot formation<br />

around Fairway Rock, located in the strait, is presented. Cross-sectional profiles<br />

of Fairway Rock and the relief of the icefoot are given along with theoretical<br />

analyses of the possible forces active during icefoot formation. It is shown that<br />

the ice cover most likely fails in flexure as opposed to crushing or buckling, as the<br />

former requires less force. Field observations reveal that the Fairway Rock icefoot<br />

is massive, with ridges up to 15 m high, a seaward face only 20' from vertical, and<br />

interior ridge slopes averaging 33'. The icefoot is believed to be grounded and its<br />

width ranges from less than 10 to over 100 meters.<br />

Introduction<br />

In the design of offshore structures to be placed in arctic waters, major consideration<br />

is being given to determining the loads developed during ice failure<br />

against a structure. This can result in the creation of an ice rubble field through<br />

which forces can be transmitted to the structure during subsequent sea ice movement.<br />

The phenomena of ice pile-up and override are also being considered.<br />

A variety of analytical and model studies have been made to investigate ice<br />

forces on offshore structures, but they are inconclusive as their results have not<br />

been verified by field measurements. This paper presents the results of a study of<br />

the general configuration of ice rubble around Fairway Rock, Alaska. The purpose of<br />

the investigation was to acquire data on the morphology of the sea ice rubble surrounding<br />

the rock that would be applicable to offshore structures in general and<br />

particularly to those placed in deep water. Estimates of the relative force levels<br />

required to form the observed rubble are also given. The surface relief across a<br />

number of sea ice rubble fields formed in the shear zone along the west side of<br />

Prince of Wales Shoal is also described for the purpose of documenting the size of<br />

ice features which may impact offshore structures placed in the northern Bering Sea.<br />

Bering Strait<br />

The Bering Strait is 85 km wide and has an irregular bottom, with a depth of 52<br />

m near the western side and about 60 m near the eastern side. On the western side of<br />

the strait is the bold topography of Cape Dezhneva and on the eastern side is the<br />

formidable landscape of Cape Prince of Wales. Winds in the strait tend to be<br />

funneled and accelerated in northerly or southerly directions by these headlands.<br />

Within the Bering Strait are three islands (Figure 1): Little and Big Diomede<br />

Islands and Fairway Rock. Fairway Rock, situated about 24 km west-southwest of Cape<br />

Prince of Wales, Alaska, is a 350-m-diameter igneous rock that rises almost vertically<br />

out of 50-m-deep water (Bloom, personal <strong>com</strong>munication) to a height of about 165<br />

USA<br />

985


Figure 2. Ice conditions in Bering Strait on 7 March 1973. Note the splitting of an<br />

ice floe moving past Fairway Rock, and the turbulent wind wakes downstream<br />

of various landforms.<br />

Coachman and Aagaard [3] show that transport through the Bering Strait is well<br />

correlated with regional east-west atmospheric pressure differences: "When a strong<br />

E-W pressure gradient lies over the strait and extends in a north-south direction<br />

from the Chukchi Sea to the central Bering Sea, entirely crossing the northern Bering<br />

Shelf, extensive northerly winds move water southward off the shelf . This produces a<br />

sea-level slope down to the south which, together with the northerly winds, drives<br />

southward transport. The atmospheric pressure pattern causing this condition is always<br />

a str ong low located a considerable distance southeast of the strait (e.g. over<br />

Kodiak) together with the Siberian high being centered some distance west of the<br />

strait."<br />

Movement of ice southward through the Bering Strait is therefore driven by<br />

northerly winds and southerly current flow. The resulting coupling of the wind and<br />

current produces drag forces on the ice which exceed its arching strength. Thus the<br />

987


Seo<br />

Figure 3. Area map of the Bering Strait .<br />

Dashed lines indicate zone of<br />

major ice stream which occurs<br />

during large southern ice drift .<br />

Stippled area is zone of maximum<br />

drift.<br />

/<br />

current produces drag forces on the ice<br />

which exceed its arching strength.<br />

Thus the arch or jammed-up ice bridging<br />

the strait fails and moves rapidly<br />

southward (Figure 2).<br />

Under strong driving forces acting<br />

over a prolonged period of time, sea<br />

ice in a belt 100 or more km wide, extending<br />

from the strait northward along<br />

the east side of the Chukchi Sea to Pt<br />

Barrow, gradually moves southward (Figure<br />

3). This movement may eventually<br />

bring multi-year ice floes into the<br />

northern Bering Sea.<br />

Once the Bering Strait ice arch<br />

collapses , in excess of 60,000 km2 of<br />

sea ice from the Chukchi Sea can move<br />

southward at high velocities in a relatively<br />

short period of time [1]. Fairway<br />

Rock is situated in the center of<br />

the major ice floe stream through the<br />

strait. As a result the rock is frequently<br />

impacted by either southerlyor<br />

northerly-moving ice floes throughout<br />

the winter season.<br />

Field Reconnaissance<br />

On 26 April 1980 a reconnaissance<br />

of Fairway Rock was made. Low cloud<br />

cover, high winds and associated turbulence<br />

around the rock caused the small<br />

aircraft to be thrown around and prevented<br />

a close-up inspection of the<br />

icefoot . Nevertheless, the photos taken show an impressive accumulation of pressured<br />

ice forming the icefoot, particularly on the north and south sides. At the time of<br />

the reconnaissance, open water surrounded the island. Off to the south significant<br />

open water and a diffused pack were noted, whereas to the north the pack ice could be<br />

988<br />

Figure 4. South side of Fairway Rock .


Figure 5. North side of Fairway Rock.<br />

Figure 6. East side of Fairway Rock. Dark objects on<br />

top of island are large propane gas tanks<br />

and a generator installation.<br />

seen in an apparent holding line between the Diomede Islands and Cape Prince of<br />

Wales . The northerly winds, while high, were not strong enough to drive the pack ice<br />

south against the prevailing southerly current .<br />

A view of the south side of the rock is shown in Figure 4. Note the steep face<br />

of the icefoot and the slope of the top of the island. The northern half of the island<br />

is seen caked with snow, and along the base the rock surface is covered with a<br />

layer of glaze ice (Figure 5). In this photo the large rock talus area is visible,<br />

as is the shear face of the massive icefoot. An east view of the rock (Figure 6)<br />

shows portions of the vertical rock face at sea level and thick icing accumulations<br />

covering the rock to an apparent height in excess of 40 m above sea level.<br />

989


Figure 7. Aerial view of Fairway Rock. Elevation profiles were made, from<br />

stereographic photo analysis, along the lines shown. North is to the<br />

right .<br />

Aerial photography of Fairway Rock was obtained on 14 March 1980. An aerial<br />

view of Fairway Rock is shown in Figure 7. The ridge systems <strong>com</strong>posing the southern<br />

icefoot are quite apparent due to the favorable sun angle . The general bluntness of<br />

the icefoot is striking, as is the narrow width of the icefoot on the west and southeast<br />

sides of the rock. In these areas the rock wall seems to drop vertically into<br />

the sea. Unlike the view of the rock shown in Figure 4 the south face of the rock at<br />

this time was caked with snow and/or glaze ice . The top of the rock is clearly windswept;<br />

visible are shadows from a cluster of propane tanks and a small generator used<br />

to power salinity, temperature, conductivity and current instruments on the sea<br />

bottom (Bloom, personal correspondence) .<br />

990


1<br />

/1<br />

/1<br />

/1<br />

/ I<br />

/ I<br />

I I<br />

Figure 14. Geometry of wedge-shaped ice sheet with idealized<br />

load distribution parameters. Note the distribution<br />

of the stress 0xx along a line at x distance from<br />

the apex.<br />

Table I. Buckling pressure for different values of wedge-angle<br />

a. R/Lp and ice thickness h.<br />

R/L P<br />

h L 0.1 1.0 10<br />

P<br />

Buckling pressure<br />

m (in. ) m ( ft) MPa ( psi) MPa ( psi) MPa<br />

a = 30·<br />

0.3 ( 11.8) 3.98 (13) 4.84 (702) 1.01 (146) 0.58<br />

0.5 (19.7) 5.85 (19) 6.45 (906) 1.30 (189) 0.74<br />

1.0 (39.4) 9.81 (32) 8.84 (1281) 1.85 (268) 1.05<br />

1.5 (59) 13.30 (44) 10.82 (1570) 2.26 (328) 1.29<br />

a = 90 0<br />

0.3 (11.8) 3.98 (13) 5.78 (838) 1.09 (158) 0.53<br />

0.5 (19.7) 5.83 (19) 7.46 (1082) 1.41 (204) 0.74<br />

1.0 (39.4) 9.81 (32) 10.55 (1529) 1.99 (289) 1.05<br />

1.5 (59) 13.30 (44) 12.92 ( 1873) 2.44 (354) 1.28<br />

a = 180 0<br />

0.3 (11.8) 3.98 (13) 9.23 (1339) 1.36 (198) 0.51<br />

0.5 (19.7) 5.83 (19) 11.92 (1729) 1.76 (255) 0.66<br />

1.0 (39.4) 9.81 (32) 16.86 (2445) 2.49 (361) 0.93<br />

1.5 (59.0) 13.30 (44) 20.64 (2995) 3.05 (442) 1.14<br />

996<br />

( psi)<br />

(84)<br />

(108)<br />

(153)<br />

( 187)<br />

(77)<br />

(107)<br />

(152)<br />

(186)<br />

(74)<br />

(96)<br />

(135)<br />

(166)


to its effective diameter. In order for this to be so the submarine slope needs to<br />

be relatively steep. At Fairway Rock it is reasonable to assume that the shallowest<br />

submarine slope was at or near the angle of repose of the rock talus.<br />

Fairway Rock appeared encrusted with a thick layer of glaze ice that extended up<br />

to 40 m above sea level. For offshore structures placed in these waters this phenomenon<br />

needs to be considered from a loading as well as an operational hindrance<br />

standpoint.<br />

The force levels given here are only estimates based upon assumed ice data and<br />

geometry considerations. Larger forces are possible at local contact pOints where<br />

stress is concentrated. The calculated effective pressures due to crushing and<br />

flexural failure of ice 1.5 m thick against Fairway Rock were estimated to be 3000<br />

kPa (435 psi) and 414 kPa (60 psi), respectively. Since the ice cover in the<br />

vicinity of Fairway Rock is believed to <strong>com</strong>prise floes not sufficiently confined to<br />

move against the rock and crush against it over its entire width, these effective<br />

pressures are considered to be high. Out of the two possible modes of failure, the<br />

ice cover would most likely fail in flexure as opposed to crushing, since it requires<br />

less force.<br />

We have presented expressions to show that buckling is also entirely possible at<br />

areas of local sheet ice contact, even when the far-field ice pressure is relatively<br />

low. Indeed, we envision that all three failure modes will occur randomly during ice<br />

failure against a large structure or during ice pack deformation.<br />

Acknowledgments<br />

Support for this study was provided by Gulf Oil Canada Resources Incorporated<br />

and in part by the Bureau of Land Management/National Oceanic and Atmospheric<br />

Administration's Alaska Outer Continental Shelf Environmental Assessment Program.<br />

References<br />

1. Ahlnas, K. and G. Wendler (1979) Sea-ice observations by satellite in the<br />

Bering, Chukchi and Beaufort Seas, <strong>Proceedings</strong> of the Fifth International<br />

Conference on Port and Ocean Engineering Under Arctic Conditions, Norwegian<br />

Institute of Technology, Trondheim, Norway.<br />

2. Bloom, G.L. and D.E. McDougal (1967) Bering Strait unattended oceanographic<br />

telemetry system utilizing a <strong>com</strong>mercial LCC-25 strontium 90 generator, The<br />

New Thrust Seaward, Transactions of the Third Annual Marine Technological<br />

Society Conference and Exhibit, 5-7 June, San Diego, Cal., Marine<br />

Technological Society, Washington, D.C.<br />

3. Coachman, L.K. and K. Aagaard (1979) Re-evaluation of water transports in the<br />

vicinity of Bering Strait, preprint of paper to appear in Bering Sea<br />

Symposium, Hood, Ed.<br />

4. Croasdale, K.R. (1980) Ice forces on fixed, rigid structures, Working group on<br />

ice forces on structures: A state-of-the-art report, T. Carstens, Ed.,<br />

CRREL Special Report 80-26.<br />

5. Kovacs, A. and M. Mellor (1974) Sea ice morphology and ice as a geologic<br />

agent in the southern Beaufort Sea, The Coast and Shelf of the Beaufort<br />

Sea, Reed and Sater, Eds., Arctic Institute of North America, Arlington,<br />

Va.<br />

6. Kovacs, A. and D.S. Sodhi (1980) Shore ice pile-up and ride-up, field<br />

observations, models, theoretical analyses, Cold Regions Science and<br />

Technology, Vol. 2.<br />

7. Kry, P.R. (1977) Ice rubble fields in the vicinity of artificial islands,<br />

<strong>Proceedings</strong> of the Fourth International Conference on Port and Ocean<br />

999


Engineering Under Arctic Conditions, Memorial University of Newfoundland,<br />

St. John's, Newfoundland, Canada.<br />

8. Michel, B. (1970) Ice pressure on engineering structures, CRREL Monograph<br />

III-BIb.<br />

9. Michel, B. and N. Toussaint (1976) Mechanism and theory of indentation of ice<br />

plates, Symposium on Applied Glaciology, Cambridge, England, Journal of<br />

Glaciology, Vol. 19, No. 81.<br />

10. Parmerter, R.R. and T.D. Coon (1973) Model of pressure ridge formation in<br />

sea ice, Journal of Geophysical Research, Vol. 77, No. 33.<br />

11. Shumway, G., D.G. More and G.B. Dowling (1964) Fairway Rock in Bering<br />

Strait, Papers in Marine Geology, Shepard Commemorative Volume, R.L.<br />

Miller, Ed., Macmillan.<br />

12. Sodhi, D.S. (1979) Buckling analysis of a wedge-shaped floating ice sheet,<br />

Fifth International Conference on Port and Ocean Engineering Under Arctic<br />

Conditions, Norwegian Institute of Technology, Trondheim, Norway.<br />

13. Sodhi, D.S. and H.E. Hamza (1977) Buckling analysis of a semi-infinite ice<br />

sheet, Fourth International Conference on Port and Ocean Engineering under<br />

Arctic Conditions, Memorial University of Newfoundland, St. John's,<br />

Newfoundland, Sept 26-30, 1977.<br />

14. Wang, Y.S. (1978) Buckling analysis of a semi-infinite ice sheet moving against<br />

cylindrical structures, International Association of Hydraulic Research<br />

Symposium on Ice Problems, Lulea, Sweden.<br />

15. Wang, Y.S. (1978) Buckling of a half ice sheet against a cylinder, <strong>Proceedings</strong>,<br />

ASCE Journal of Engineering Mechanics Division, EMS.<br />

1000


Stuart D. Smith<br />

and<br />

Erik G. Banke*<br />

ABSTRACT<br />

A NUMERICAL MODEL OF ICEBERG DRIFT<br />

Atlantic Oceanographic Laboratory<br />

Bedford Institute of Oceanography<br />

Dartmouth, Nova Scotia B2Y 4A2<br />

The movement of icebergs under the influence of winds and currents<br />

has been hindcast using a simple numerical model. Air and water drag<br />

coefficients have been adjusted to give a best fit to the observed drift<br />

in five of seven cases investigated, while in two other cases the observed<br />

winds and currents cannot explain the observed track.<br />

INTRODUCTION<br />

Exploration activity in iceberg infested waters off the east coast<br />

of Canada has brought new urgency to problems of tracking and predicting<br />

iceberg movement. It will eventually be necessary to predict these<br />

motions in the vicinity of a rig or structure over distances of a few<br />

tens of kilometers and time scales of about one day.<br />

In the present study we attempt to hindcast the motion of icebergs<br />

using data which have been taken on a routine basis. The purposes of<br />

this study are to determine whether, or in what circumstances, these<br />

data are adequate to describe the dynamics of iceberg motion, and to<br />

infer what additional data may be necessary to better describe the<br />

dynamics. In addition we are able to examine the relative influence of<br />

winds and currents, and will in a future paper use this model to predict<br />

what influence towing forces would have had on the modelled tracks.<br />

*Nowat: Martec Ltd., 1526 Dresden Row, Halifax, N.S.<br />

Canada<br />

1001


Coriolis Force<br />

This fictitious force allows for the rotation of the earth, which<br />

makes a floating object appear to drift in a circular path in the<br />

absence of drag or other applied forces. Corio lis force also acts on<br />

the water in which the iceberg floats. The water surface is assumed to<br />

slope in geostrophic balance with the Coriolis force, so that an iceberg<br />

moving with the water would experience no net deflection. Only the<br />

velocity of the iceberg relative to the water<br />

V<br />

l'<br />

+ + +<br />

V - W = -w l'<br />

is used to <strong>com</strong>pute the Coriolis force<br />

F = Mf X V<br />

C l'<br />

where the Coriolis vector<br />

If I = 2Qsin'"<br />

is directed vertically upward, Q = 7.27 X 10- 5 rn s-l is the earth's<br />

rate of rotation, and", is the latitude. No adjustments to the esti­<br />

mated mass M are made in optimizing the fit of the model to the observed<br />

track, but a proportional increase (decrease) of both air and water drag<br />

coefficients together is equivalent to a proportional decrease (increase)<br />

in the mass.<br />

MODELLING PROCEDURE<br />

For each track to be modelled the initial wind, current, position,<br />

and velocity are supplied, and a description of the iceberg in terms of<br />

sail and keel area, mass, and assumed air and water drag coefficients.<br />

At six second intervals a <strong>com</strong>puted force balance is used to update the<br />

velocity and position. New data are read in at regular hourly intervals<br />

in the examples to be shown, but the model allows these data to be<br />

specified at irregular intervals in 10 minute steps.<br />

The observed track is plotted starting at time zero as a line with<br />

a square symbol at each observed point and a heavier symbol at every<br />

twelfth point. The <strong>com</strong>puted track is plotted in 10 minute line segments<br />

with a symbol every hour and a heavier symbol every 12 hours. Any<br />

number of models of the same track can be plotted on a page, each being<br />

identified by a different symbol, and a number of tracks with different<br />

(6)<br />

(7)<br />

(8)<br />

1003


choices of C a and C w can be <strong>com</strong>pared subjectively. At each observed<br />

point (i.e. hourly), the magnitude of the vector difference between<br />

observed and <strong>com</strong>puted drift since the previous observation is used to<br />

find an rms error over the entire track. Selecting C a and C w for<br />

minimum rms errors in the individual (hourly) drift segments is an<br />

objective method which must be checked to ensure that it corresponds to<br />

a reasonable cumulative drift track.<br />

MODELLING OF OBSERVED ICEBERG TRACKS<br />

A number of iceberg drift observations have been made available to<br />

us (see Acknowledgements). About half of the tracks supplied were not<br />

used because the data were in<strong>com</strong>plete or irregular in time. In six<br />

cases we had estimates of the height, width, and mass of the iceberg<br />

(Table 1), a sketch of its above-water shape, and an hourly log of<br />

iceberg range and bearing, and wind and current speed and direction,<br />

measured at the drillship Pelican. Currents at 50 metre depth were used<br />

when available, since we feel that currents at the only other available<br />

depth (IS m) may be subject to influences from the ship's hull. In<br />

cases where the draft was not known the keel area was estimated to be<br />

four times the sail area.<br />

The variability of current between the drill ship and an iceberg at<br />

2 to 20 km distance is obviously a major source of uncertainty, and<br />

success of our model depends on the currents being uniform over the<br />

separation distance. An array of current meters would be required to<br />

properly define the currents at the iceberg location. From the level of<br />

success achievable in a number of modelling attempts in a particular<br />

area, we hope to be able to estimate how dense such an array would have<br />

to be.<br />

Iceberg K007<br />

Although this was the last of the group chronologically, its track<br />

is especially interesting and will be discussed in more detail than the<br />

others. This was the largest of the group, and drifted close to the<br />

ship so that the measured currents should be representative of those<br />

acting on the iceberg.<br />

1005


1008<br />

We concentrated on modelling a 54 hour segment of the track which<br />

is nearest the ship. The iceberg was grounded for 19 hours of this<br />

period, which we modelled kinematically (not dynamically) by simply<br />

setting the velocity to zero during the appropriate period. An optimum<br />

track with rms error 0.5 km was obtained with C = 0.55 and C - 0.57<br />

a w<br />

(Figure 1).<br />

The effects of variations in the drag coefficients are illustrated<br />

in Figure 2. In the upper group, the air and water drag coefficients<br />

are both doubled and then both halved with the track be<strong>com</strong>ing longer and<br />

straighter, or shorter and curlier. In the lower group, the coeffi­<br />

cients are each (in turn) incremented up and down by 0.1. Increasing C w<br />

moves the end point after 54 hours in a southeasterly direction, while<br />

increasing C a moves it in a west-southwesterly direction. Figure 3<br />

shows that the wind drift and the current drift (with winds and currents<br />

in turn set to zero) are of similar importance in determining the total<br />

drift.<br />

Iceberg<br />

F018<br />

F025<br />

K016A<br />

K016B<br />

K024<br />

S012<br />

K007<br />

TABLE 2. PRELIMINARY RESULTS OF MODELLING ICEBERG TRACKS<br />

Best rms error Length of Track<br />

C<br />

a<br />

km in 1 hr hr.<br />

km<br />

0.05 0.55 0.51 22<br />

9<br />

0.1 0.6 0.37 17<br />

7<br />

10<br />

11<br />

t t<br />

36<br />

40<br />

t t<br />

18<br />

25<br />

1.0 0.3 1.05 23<br />

33<br />

0.55 0.57 0.51 54*<br />

35<br />

* Aground for 19 hours of this period<br />

t Unable to model this track<br />

Distance (km)<br />

due to<br />

current wind<br />

8<br />

5<br />

9<br />

13<br />

10<br />

16<br />

20<br />

1<br />

2<br />

2<br />

t<br />

t<br />

17<br />

15<br />

Range<br />

km<br />

8-17<br />

15-20<br />

18-25<br />

18<br />

15-19<br />

22-28<br />

2-15<br />

Table 2 lists the results to date. In general the current appears<br />

to have a slightly larger effect than the wind on the drift. The track<br />

of iceberg K016 was split into a 10-hour period which was modelled<br />

(K016A) and a 36-hour period during which the iceberg continued to drift


1010<br />

in a southerly direction, while the currents were slow and circled in a<br />

clockwise direction, and the winds were light. The track of iceberg<br />

K024 similarly could not be modelled because the observed drift was to<br />

the southeast while the modelled track would give a slow, clockwise<br />

motion driven mainly by the currents.<br />

DISCUSSION<br />

The results given here are preliminary and development of the model<br />

is continuing. Even with sparse data we are able to achieve some success<br />

by using an adaptive modelling approach in which the observed track is<br />

used to over<strong>com</strong>e deficiencies in knowledge of the size and drag characteris­<br />

tics of an iceberg. In a minority of cases the measured currents and<br />

winds appear not to be related to the drift tracks.<br />

Many hundreds of iceberg tracks have been observed from drill ships<br />

and rigs off the coast of Labrador and elsewhere, and we plan to continue<br />

testing and improving our model as more of these data be<strong>com</strong>e available.<br />

In particular, if currents at a number of locations and depths are measured,<br />

an interpolation to the iceberg location will be attempted. Separate<br />

water drag calculations for several depth ranges (e.g. Mountain, 1980)<br />

would then be possible.<br />

We are presently experimenting with the addition of towing forces<br />

to ,the successfully modelled tracks. This will allow us to estimate how<br />

much deflection could have been achieved if a particular towing force<br />

had been exerted over a particular time interval. Another area of<br />

ongoing development is in automating the selection of the best drag<br />

coefficients, and in further developing objective criteria for a "best"<br />

fit.<br />

Forecasting<br />

In order to use this model for forecasting, it would be necessary<br />

to forecast the currents for several hours. While tidal <strong>com</strong>ponents of<br />

current can be identified and forecast accurately, and forecasts of<br />

wind-driven (Ekman) currents can be attempted, variations associated<br />

with large-scale ocean circulation cannot be forecast by a local model.<br />

Wind forecasts are routinely available, and can be improved if a number<br />

of high-quality offshore surface and radiosonde weather observations are<br />

added to the regular meteorological network. Site-specific forecast


services are based on the large scale numerical forecasts, and the<br />

offshore weather data could be more fully used to feed and tune-up these<br />

numerical forecasts.<br />

Development of the present model into a forecast mode can proceed<br />

in two steps: (1) Using only part of the track to optimize the drag<br />

coefficients, and then using observed winds and currents to hindcast the<br />

rest of the track with these coefficients, and (2) Replacing the winds<br />

and currents with forecasts based on data available at the time when<br />

the forecasting mode is started. This approach would allow separation<br />

of errors into those associated with iceberg dynamics and those due to<br />

wind and current forecasting errors.<br />

While it may be some time before iceberg drift models be<strong>com</strong>e<br />

operationally useful, the present practice is to tow icebergs if they<br />

appear to endanger an operation, based on extrapolation of the observed<br />

drift rate. It should be possible to develop an adaptive model to help<br />

in decision making by giving an estimate the effects of anticipated<br />

changes in winds and currents, and the possible effectiveness of towing.<br />

ACKNOWLEDGEMENTS<br />

The data were collected by MacLaren Marex (now Canplan) Ltd., under<br />

contract to Total Eastcan Ltd., and were made available to us by P.E.R.<br />

Vandall of Resource Management Branch, Dept. of Energy, Mines and Re­<br />

sources, Ottawa.<br />

REFERENCE<br />

Mountain, D.G., 1980: On predicting iceberg drift. Cold Regions<br />

Science and Technology, !, 273-282.<br />

1011


ICEBERG SCOUR STUDIES IN MEDIUM DENSE SANDS<br />

T. R. Chari Faculty of Engineering & Applied Science<br />

Memorial University of Newfoundland<br />

H. P. Green St. John's, Newfoundland Canada<br />

ABSTRACT<br />

INTRODUCTION<br />

The problem of iceberg scours on Canada's east coast is a major<br />

hazard in the extraction of the offshore hydrocarbon resources. Various<br />

production systems which take into account the severe environmental fac­<br />

tors such as the heavy seas, ice and icebergs are under consideration<br />

for the Hibernia field on the Grand Banks. In any system, all seafloor<br />

structures are to be located below the zone of iceberg scours. However,<br />

the estimation of the maximum scour depths is still an aspect of the<br />

problem not fully understood. A model for iceberg scouring in clays has<br />

been suggested earlier and is now modified to include cohesionless<br />

soils. Laboratory tests were conducted with a 50 cm wide model, using<br />

medium dense sand as the representative seabed material.<br />

these experiments are discussed.<br />

Results of<br />

The problem of seabed scouring by icebergs and the consequent threat to the pipe­<br />

lines and buried installations on the Canadian eastern seaboard is well recognized.<br />

Scours as deep as 6.5 m have been measured (Harris and Jollymore, 1974) in some<br />

locations. Geological surveys of the different oceans have revealed scour-like fea­<br />

tures in various locations where icebergs and ice are no longer present day phenomena.<br />

The Northeast Atlantic west of the British Isles is a typical example, where furrows<br />

up to about 25 m wide have been observed (Belderson et aI, 1973) with depths of 5 m<br />

and lengths of 2 \an in waters of 140 m to 500 m. Such observations have lead to some<br />

1012


of safety of 1.5 to 3 usually adopted in s t ructural and Geotechnical engineering<br />

designs, this difference is not considered to be very significant.<br />

LABORATORY MODEL TESTS<br />

The significance of the laboratory tests, their interpretation and relevance to<br />

the mathematical model have been discussed elsewhere in detail. (Chari 1975, 1980).<br />

Results of tests in a tilting flume using a cohesive sediment as the representative<br />

seafloor soil were also reported . Laboratory experiments on iceberg scouring are<br />

designed to verify only the nature and type of soil resistance forces. These tests do<br />

not duplicate the entire mathematical model as given by Eqns . [IJ and [2]. The mathe­<br />

matical model consists of different <strong>com</strong>ponents, the kinetic energy of the moving ice­<br />

berg, the hydrodynamic drag, the soil resistance forces and the principle of energy<br />

balance. The problems of scale modelling all these properties, particularly the soil<br />

strength, precludes the possibility of duplicating the entire scour phenomenon in the<br />

laboratory . Further, all the concepts except the nature and type of soil resistance<br />

are well established principles in Applied Mathematics and in engineering requiring no<br />

further experimental verification . Thus the laboratory experiments have been designed<br />

to measure the forces and pressures on a physical model and also inside the soil<br />

medium when the idealized model is towed at constant speeds into a sloping bed of the<br />

sediment. Laboratory measurements attempt to justify the right hand side of Eqns .<br />

[1 J and [2 J which then validates the entire mathematical model. Recent tests in the<br />

FIG. 2: LABORATORY TEST FACILITY<br />

1015


parison of the theoretical and experimental values for the tests in sand. It is ob­<br />

served that the measured soil resistance is higher than the theoretical values for the<br />

cohesionless soils as well. However, the maximum difference is in the order of 10% as<br />

against 30% for the clays reported earlier (Chari, 1980). A similar phenomenon has<br />

been reported by Krause (1974) for experiments in loose sands. Consistent with the<br />

above observations, it was postulated (Chari, 1980) that there would be a movement of<br />

the soil ahead and below the actual limits of the scour. This was demonstrated by<br />

placing pressure cells inside the soil mass which recorded pressures when the iceberg<br />

model passed above. To confirm these observations further, in the recent tests with<br />

the cohesionless soil, a model pipeline was instrumented and placed in the soil at<br />

different locations relative to the scour boundary (Fig. 4). A set of typical results<br />

is shown in Fig. 5. These confirm that there is a movement of the soil below and<br />

ahead of a scouring iceberg. Maximum pressures of 12.5 kPa were measured in these<br />

experiments. Further work is in progress to delineate the zone of soil movement and<br />

also to evaluate the scale effects.<br />

CONCLUSIONS<br />

The model for iceberg scouring in soft clayey sediments was extended to fric­<br />

tional soils. There are some problems in the precise evaluation of the type of fric­<br />

tional forces between the Boil and iceberg. However, preliminary <strong>com</strong>putations show<br />

that for an iceberg of 10 x 10 9 kg the scour depth would be overestimated in the<br />

FIG. 4: PIPELINE MODEL EMBEDDED IN SOIL<br />

1017


John D. Miller<br />

Senior Environmental<br />

Analyst - Ice<br />

ABSTRACT:<br />

A SENSITIVITY ANALYSIS OF A SIMPLE MODEL OF<br />

SEASONAL SEA ICE GROWTH<br />

Petro-Canada<br />

Box 2844<br />

Calgary, Alberta<br />

T2P 3E3<br />

Canada<br />

Use is made of a simple model of seasonal sea ice growth to evaluate the sensitivity<br />

of ice growth to variations in the climatological forcing functions. The<br />

model is an energy conservation, equilibrium surface temperature simulation that<br />

has been found in previous studies to realistically reproduce first year sea ice<br />

growth in Arctic locations. By employing a standard test case in conjunction<br />

with varying climatological conditions the growth response of the ice to the new<br />

conditions is evaluated. A series of simulations to evaluate the effects of<br />

changes in air temperature, wind speed, in<strong>com</strong>ing solar radiation, snow density<br />

and snow depth are performed and the results reported upon.<br />

INTRODUCTION<br />

With the increasing development of the Arctic regions the need has be<strong>com</strong>e apparent<br />

for a simple yet rigorous method to estimate the growth response of a sea<br />

ice cover and the associated energy fluxes. As well, a knowledge of the thickness,<br />

properties and coverage of young ice is important in studies of climatic<br />

change, regional energy balance, ice dynamics, ice forecasting, ice engineering,<br />

remote sensing response, petroleum extraction and marine transportation. To this<br />

end a simple model of heat transport through young sea ice coupled with the energy<br />

exchange mechanisms in the lower atmosphere has been developed to estimate ice<br />

growth and decay in association with the ice surface micrometeorology. With this<br />

model the sensitivity of ice growth to changing environmental conditions may be<br />

evaluated.<br />

MODEL DEVELOPMENT<br />

The use of climatological models to predict sea ice growth offers a variety of<br />

advantages over empirical or analytical techniques. Empirical techniques are<br />

limited in that a host of energy exchanges are subsumed and expressed in terms of<br />

a surrogate variable or variables and a relatively simple functional relationship.<br />

This simplicity defies an understanding of process interaction resulting<br />

in a 'black box' approach which makes it impossible to investigate, or masks the<br />

effect of, changes in other related parameters. Analytical approaches include<br />

more effects but solutions are often limited to cases where relatively simple<br />

boundary conditions exist.<br />

1020


Maykut and Untersteiner (1969) presented a <strong>com</strong>prehensive one dimensional thermodynamic<br />

model of sea ice in which the turbulent energy fluxes, radiation fluxes,<br />

ocean heat fluxes, snow accumulation and density, surface albedo and ice salinity<br />

are treated as time dependent model inputs. With these the energy balance is calculated<br />

as a function of an effective surface temperature and the ice mass changes<br />

(accretion/ablation) calculated at the ice boundaries. A pioneering stage in the<br />

modelling of ocean-ice-atmosphere interaction, it remains a most <strong>com</strong>prehensive<br />

treatment of ice thermal behaviour.<br />

Since this work others have developed models which attempt to interrelate climatological<br />

processes to the ice and its growth. Notable works include those of<br />

Goddard (1974), Semtner (1976), Washington et al (1976) and Maykut (1978).<br />

For this work a simulation of climatological fluxes and ice growth employing energy<br />

conservation methodology was adopted. The <strong>com</strong>plete development of the model is<br />

presented in Miller (1979, 1980) and is summarized here. The model attempts to<br />

<strong>com</strong>bine the synoptic scale climatic variables with site characteristics to evaluate<br />

the surface energy balance. This balance is given by the energy conservation<br />

equation (equation 1) with the magnitude and direction of its <strong>com</strong>ponents resulting<br />

from the thermal, radiative and aerodynamic properties of the site.<br />

Q* + FH + FE + FI + FA = 0 (1)<br />

where Q* net radiation flux<br />

FH sensible heat flux<br />

FE evaporative heat flux<br />

FI conductive heat flux<br />

FA energy flux due to ablation<br />

The <strong>com</strong>ponents of the net radiation flux can be expressed as the sum of four radiation<br />

terms: in<strong>com</strong>ing and outgoing (reflected) shortwave radiation and the in<strong>com</strong>ing<br />

and outgoing longwave radiation. In the model the in<strong>com</strong>ing shortwave radiation<br />

is required as a specified input. The outgoing radiation is expressed, in<br />

the presence of a snow cover, as:<br />

Kt = (Hi (2)<br />

where a = surface albedo<br />

Under snowfree conditions penetration of the shortwave radiation occurs into the<br />

ice. To allow for this phenomenon the approach of Maykut and Untersteiner (1969) is<br />

employed in which a predetermined percentage of the net shortwave radiation is<br />

assigned to penetration of the ice and the remaining portion used to determine<br />

the surface energy balance. This percentage is fixed during the snow free period.<br />

The net shortwave (K*) is then calculated as:<br />

K* = (1-a) (1-i) Ki (3)<br />

where i = penetrating fraction<br />

In<strong>com</strong>ing longwave radiation is estimated using an empirical relation developed<br />

by Idso and Jackson (1969).<br />

Lj= fUT 4(1-0.261 exp (-7.77 x 10 -4(T -273)2)<br />

a a<br />

(4)<br />

1021


The ice thermal conductivity exhibits a temperature and salinity dependency as<br />

defined by Untersteiner's (1961) functional relation. Snow thermal conductivity<br />

is determined from the snow density ( P s )' using Van Dusen's (1927) equation.<br />

The average ice salinity (parts per thousand) is estimated on the basis of ice<br />

thickness using the empirical relationships obtained by Cox and Weeks (1974).<br />

In accordance with the observations of Thorpe et al (1973) and Banke et al (1976)<br />

who found that CD»CH»C E , in the model C H is assigned the value 1.1 x 10- 3 with<br />

C E set to 0.6 x 10- 3 in accordance with their observations.<br />

Air density is fixed at 1.32 kg/m 3 , the specific heat capacity of air at 1010 J<br />

kg- 1 ·C-l and the latent heat of vaporization at 2.533 x 10 6 J/kg. Ice density<br />

is assumed to be 920 kg/m 3 , the latent heat of formation fixed at 2.72 x 105 J /kg<br />

while the latent heat for melting of snow and ice is taken as 3.344 x 105 J/kg.<br />

The ocean temperature beneath the ice is taken as the salinity determined freezing<br />

point, a feature supported by observational data of Lake and Lewis (1971);<br />

Doherty and Kester's (1974) relationship between freezing point and salinity is<br />

used.<br />

MODEL RESULTS<br />

The model has been previously tested against ice growth data at three Arctic sites<br />

which provide a significant range of climatic environments. Nine years of simulations<br />

at Eureka N.W.T. (80 0 00'N, 85 0 46'W), Frobisher Bay, N.W.T. (63 0 45', 68°33'W)<br />

and Resolute N.W.T. (74 0 43'N, 94 0 59'W) are presented and evaluated in Miller (1979).<br />

It was found that the model accurately reproduced the pattern and magnitude of ice<br />

growth under a variety of climatological conditions.<br />

The test site chosen for the sensitivity analysis was Resolute N.W.T. which features<br />

a relatively long period of ice record (since freeze-up 1955), an average<br />

ice season of greater than 280 days and operates as a first order climatological<br />

station. Climatological normals were obtained from the Atmospheric Environment<br />

Service (1972, 1976) records and are summarized in Table One, while summary ice<br />

growth data were found in Richardson and Burns (1975).<br />

The sensitivity analysis was performed by first defining a normal or standard case<br />

and defining the ice growth season on a daily basis (see Figure One). Specified<br />

parameters were then changed, singly or in unison, and the ice season simulated.<br />

Changes between the standard case and the simulated case were then identified and<br />

evaluated to determine the effect of the variation of the parameter under investigation.<br />

The standard case was run using the data presented in Table One. Daily values<br />

were obtained by employing a cubic spline interpolation on the monthly mean values<br />

of temperature, in<strong>com</strong>ing solar radiation and wind speed; the snowfall was allowed<br />

to accumulate in two snowfalls per month with a density of 310 kg/m 3 • An initial<br />

ice thickness of 29 cm and starting date of October 01 (Julian day 274) were used.<br />

With these data a maximum ice thickness of 178.4 cm occurs on Julian day 173. The<br />

snow is ablated over six days while 117.8 cm of ice is lost during the next sixty<br />

three days yielding an end of season thickness of 60.6 cm and a sixty nine day<br />

melt period. Comparison with the Richardson and Burn (1975) data show a very reasonable<br />

similarity between the simulated standard condition and the statistically<br />

derived normals.<br />

1025


7 Lake, R.A. and E.L. Lewis 1971. The microclimate beneath growing<br />

sea ice. Proc. I.A.H.R. Conf. (T. Karlsson, Ed.) Reykjavick, May 10-13,<br />

1971, N.R.C., 241-244.<br />

8 Maykut, G.A. 1978. Energy exchange over young sea ice in the<br />

central arctic. J. Geophys. Res., 83(C7): 3646-3658.<br />

9 Maykut, G.A. and N. Untersteiner 1969. Numerical prediction of the<br />

thermodynamic response of arctic sea ice to environmental changes, Rand<br />

Memorandum RM-6093-PR, Nov. 1969, 173 p.<br />

10 Miller, J.D. 1979. An equilibrium surface temperature climate model<br />

applied to first year sea ice growth. Unpublished Masters Thesis, Carleton<br />

University, 182 p.<br />

11 Miller, J.D. 1980. A simple model of seasonal sea ice growth. ASME<br />

80-WA/HT-20, 8 p.<br />

12 Richardson, F.A. and B.M. Burns 1975. Ice thickness climatology for<br />

Canadian stations. A.E.S. Publication Ice 1-75, 60 p.<br />

13 Semtner Jr., A.J. 1976. A model for the thermodynamic growth of<br />

sea ice in numerical investigations of climate. J. Phys. Oceanography, 6:<br />

379-389.<br />

14 Thorpe, M.R., Banke, E.G., and S.D. Smith 1973. Eddy correlation<br />

measurements of evaporation and sensible heat flux over arctic sea ice.<br />

J. Geophys. Res., 78(18): 3573-3584.<br />

15 Untersteiner, N. 1961. On the mass and heat budget of the arctic<br />

sea ice. Arch. Meteorol. Geophys. Bioklimatol., A, 12, 151-182.<br />

16 Van Dusen, M.S. 1929. IntI. Crit. Tables, 5: 216.216.<br />

17 Washington, W.M., A.J. Semtner Jr., C. Parkinson and L. Morrison,<br />

1976. On the development of a seasonal change sea-ice model. J. Phys.<br />

Oceanogr., 6: 679-685.<br />

18<br />

(14):<br />

1030<br />

Weller, G. 1972.<br />

28-30.<br />

Radiation flux investigation. AIDJEX Bulletin


Mr. Matti Lepparanta<br />

Prof. Erkki Palosuo<br />

STUDIES OF SEA ICE RIDGING WITH A<br />

SHIP-BORNE LASER PROFILOMETER<br />

Institute of Marine Research<br />

University of Helsinki<br />

Abstract<br />

Finland<br />

Finland<br />

Distributions of the height and spacing of ridge sails have been<br />

observed with a laser profilometer mounted on the deck of ice­<br />

breakers. The laser beam is directed down to one side of the ship<br />

and it hits the surface at a distance of 15-20 m from the ship. The<br />

laser is in use in the Baltio Sea in a long-term program for mapping<br />

ridge statistics. The results show that the average sail height is<br />

typically 45-55 cm and linear ridge density 5-10 km- 1 , when the cut­<br />

off height is defined as 30 cm; both sail heights and spacings<br />

fit well the negative exponential distribution. The mass of ridged<br />

ice is estimated to be typically 0.2-0.6 times the level ice mass.<br />

The laser was used in July 1980 in the area from Svalbard to Franz<br />

Josef Land between the latitudes of 78 0 N and 82 0 N during the Ymer-80<br />

expedition. The method worked well in fipst-year ice but in heavier<br />

ice the ship's run was too uneven.<br />

1. Introduction<br />

Spatial features of sea ice ridges are presently measured mainly<br />

through linear profiling. Upper portions have been recorded with<br />

airborne lasers (e.g., [3]). In the Baltic Sea a laser profilometer<br />

has been used in the last three years from icebreakers and the<br />

1031


5. Concluding remarks<br />

Spatial distribution of sea ice ridges can be observed by using a<br />

laser profilometer aboard an icebreaker. The laser beam is directed<br />

down to one side of the ship and it hits the surface at a distance of<br />

15-20 m from the ship. The method has been tested in the Gulf of<br />

Bothnia, Baltic Sea, and taken in use in a long-term program for<br />

mapping ridge statistics. The measurements can be done while the ice­<br />

breaker is performing her routine assistance work.<br />

Aboard 16 MW icebreakers the method works well in first-year ice.<br />

This is especially the case in the marginal ice zone and subarctic<br />

seas where the level ice sheet and totally frozen layer of ridges are<br />

not thick.<br />

6. References<br />

1. Gudkovic, Z.M. & M.A. Romanov 1976: Method for calculation the<br />

distribution of ice thickness in the Arctic seas during the<br />

winter period. - In: Krutskih, B.A., Z.M. Gudkovic & A.L.<br />

Sokolov (eds.): Ice Forecasting Techniques for the Arctic<br />

Seas, pp. 1-48. New Delhi (translated from Russian) •<br />

2. Hibler, W.D.,III, W.F. Weeks & S.J. Mock 1972: Statistical aspects<br />

of sea ice ridge distributions. - J. Geophys. Res. 77:5954-70.<br />

3. Ketchum, R.D., Jr. 1971: Airborne laser profiling of the Arctic<br />

pack ice. - Remote Sensing Environ. 2:41-52.<br />

4. Kirillov, A.A. 1957: Calculation of hummockness in determining ice<br />

volume. - Probl. Arctic 2:53-58.<br />

5. Lepparanta, M. 1981a: On the structure and mechanics of pack ice<br />

in the Bothnian Bay. - Finnish Mar. Res. 248.<br />

6. -"- 1981b: Statistical features of sea ice ridging in the Gulf of<br />

Bothnia. - Styrelsen f8r Vintersj8fartsforskning, Forskningsrapport<br />

32. Helsinki.<br />

7. Makinen, E., A. Keinonen & A. Laine 1976: Ice resistance measurements<br />

with IB APU in the Baltic Sea. - Ocean Engng. 3:267-91.<br />

8. Palosuo, E. 1975: Formation and structure of ice ridges in the<br />

Baltic. - Styrelsen f8r Vintersj8fartsforskning, Forskningsrapport<br />

12. Helsinki.<br />

9. Wadhams, P. 1980: A <strong>com</strong>parison of sonar and laser profiles along<br />

corresponding tracks in the Arctic Ocean. - In: Pritchard, R.<br />

(ed.): Proc. ICSI/AIDJEX Symp. on Sea Ice Processes and<br />

Models, University of Washington.<br />

1037


CHUKCHI SEA ICE MOTION<br />

R. W. Reimer<br />

and<br />

J. C. Schedvin, Research Scientist<br />

R. S. Pritchard, Sr. Research Scientist<br />

Flow Research Company<br />

Kent, Washington 98031 U.S.A.<br />

Abstract<br />

Chukchi Sea ice motion in response to ocean currents is simulated to determine if<br />

oiled ice could be transported from the Alaskan North Slope oil fields over 1000 km<br />

southward into the Bering Sea. This is the first such detailed simulation of the<br />

ice behavior in this region. Reversals of the typically northward ocean currents<br />

through the Bering Strait and Chukchi Sea provide the dominant driving force that<br />

induces breakouts and large-scale southward ice motions. Therefore, we use an ocean<br />

current model coupled with the AIOJEX elastic-plastic model to describe ice<br />

behavior. The ocean current field is extrapolated over the Chukchi Sea from data<br />

collected at seven current meters during the winter of 1976-77. Since large-scale<br />

cumulative motions depend significantly on ice strength, strength is varied from<br />

zero to 10 6 N/m. For zero strength (free drift), ice is transported from Point<br />

Barrow to within 100 km of the Bering Strait; for a strength of 10 4 N/m, ice is<br />

transported from Point Barrow nearly to Cape Lisburne; for a strength of 105 N/m,<br />

only 100 km of motion is simulated; and finally, for a strength of 10 6 N/m, no<br />

motion occurs. This simulation shows that it is unlikely that oiled ice could be<br />

carried southward from Point Barrow through the Bering Strait. Even at the lowest<br />

strength limit (free drift), the oiled ice does not reach the Bering Strait during<br />

the winter of 1976-77.<br />

Acknowledgement<br />

This study was funded by the Bureau of Land Management through interagency agreement<br />

with the National Oceanic and Atmospheric Administration as part of the Outer<br />

Continental Shelf Environmental Assessment Program.<br />

1038


Introduction<br />

As part of the concern over oil spill hazards in the offshore Prudhoe Bay lease<br />

areas, consideration must be given to the possibility of spilled oil being<br />

transported through the Bering Strait and into the Bering Sea. Even though Prudhoe<br />

Bay and the Bering Strait are separated by over 1000 km, it is possible to envision<br />

a sequence of events which could lead to oiled pack ice passing through the Bering<br />

Strait. This paper addresses the likelihood of such large-scale motions in the<br />

Chukchi Sea by determining for the first time the velocity field of the Chukchi Sea<br />

ice cover during a strong southward motion. Specifically, we ask: Given that oiled<br />

ice is near Point Barrow, what is the likelihood of its being transported southward<br />

across the Chukchi Sea and through the Bering Strait? This question cannot be<br />

answered with certainty because of limitations in our knowledge of ice behavior and<br />

ocean currents in the Chukchi Sea. However, in our simulations, we determine ice<br />

trajectories for conditions that tend to maximize the southward flow of the pack<br />

ice. As a result, the model simulations provide a worst case for the transport of<br />

oiled ice through the Bering Strait.<br />

Previous work has shown that reversals of generally northward-flowing currents are<br />

the cause of large, southward ice motions [1]. Drag due to ocean currents has been<br />

identified as the major driving force for the breakout of the Chukchi Sea ice pack<br />

through the Bering Strait, with wind stress and traction at the northern boundary of<br />

the Chukchi Sea playing only secondary roles [2]. On the basis of these results, we<br />

calculate ice velocity fields in the Chukchi Sea during a time of large-scale<br />

current reversal using a realistic model of the Chukchi Sea currents to provide the<br />

current drag forces. From the ice velocity fields we are able to calculate the<br />

southward trajectory of sea ice through the course of one winter.<br />

Ice-Ocean Model<br />

In this work a mathematical model which is capable of being driven by winds and<br />

ocean currents is used to simulate ice behavior. Forces on the ice occur as a<br />

result of divergence of internal ice stress, drag on its bottom surface by ocean<br />

currents and across its top surface by winds, sea-surface tilt, and Coriolis<br />

acceleration. This allows a range of inputs to be introduced to find a range of<br />

motions. Following Reimer et al. [2], we neglect wind stress and the stress applied<br />

by the Beaufort Sea ice pack, as these were found to have only a minor influence on<br />

the southward transport of ice in the Chukchi Sea. We consider variations in ice<br />

conditions and currents to determine how each affects the ice behavior. Pack ice is<br />

modeled as an elastic-plastic material, a continuum on the scale of tens of<br />

kilometers. The ice model used is essentially the same as that developed by<br />

1039


AIDJEX [3], but strength is taken as a constant for each simulation and does not<br />

vary as the ice deforms [4].<br />

In order to develop a realistic model of the ocean currents, a study was made of<br />

available oceanographic current data. Much of this information is collected and<br />

summarized by Coachman et al. [5]. More recent data from current meter moorings are<br />

discussed by Tripp et al. [6] and Coachman and Aagaard [7]. The later data include<br />

observations from a 5-month period during the winter of 1976-77 when the Chukchi Sea<br />

was ice covered. Dur ocean current model is an extrapolation of these current meter<br />

measurements in accord with the data prior to 1975.<br />

The ocean current model is generated by dividing the Chukchi Sea into a number of<br />

generally north-south trending regions. These regions extend from the Bering Strait<br />

to an arc passing from Wrangel Island to Point Barrow. The boundaries of the<br />

regions follow the expected current paths; therefore, the northward or southward<br />

transport in each region is constant (northward and southward are taken to mean away<br />

from and toward the Bering Strait, respectively). Fig. 1 shows the boundaries of<br />

the current path regions and the locations of current meters NC-l through NC-7 in<br />

the Cape Lisburne section and NC-IO in the Bering Strait. The north-south water<br />

transport in each region is determined by multiplying the cross-sectional area of<br />

the region at the current meter by the perpendicular velocity <strong>com</strong>ponent measured by<br />

the meter. Local current magnitudes, at locations other than current meter<br />

stations, are determined from the local transport, conserved in each region, divided<br />

by the local cross-sectional area of the region. Current meter NC-IO is used to<br />

determine the transport through the eastern half of the Bering Strait. The<br />

transport through the western half of the Bering Strait is the difference between<br />

the Cape Lisburne section transport and that through the eastern half of the<br />

Strait. This allows a nonuniform flow through the Bering Strait that corresponds to<br />

reported observations.<br />

The current field constructed from the model for 4 February 1977 is shown in<br />

Fig. 2. This is the largest observed southward current event during the 1976-77<br />

experiment when the Chukchi Sea was ice covered; southward current speeds at Cape<br />

Lisburne reached values of 0.5 m/s. Using this current distribution, the ice<br />

velocity field is calculated for the entire Chukchi Sea. Separate calculations are<br />

made for four values of ice strength, p*. With zero ice strength, free drift<br />

occurs. In free drift, the ice velocity field is identical to the local current<br />

distribution shown in Fig. 2. Fig. 3 shows the ice velocity fields for p* = 10 4<br />

and 105 N/m. When p* is set to 10 6 N/m, no ice motion occurs.<br />

1040


Figure 1. Current Meter Locations.<br />

Transport Sections. and Current Path<br />

Regions<br />

1 m/s<br />

8. Strength p* = 104N/m<br />

1 m/s<br />

Figure 2. Ocean Current Field on<br />

4 February 1977. Also Free Drift<br />

Ice Velocity Field.<br />

1 m/s<br />

b. Strength p* = 10 5 N/m<br />

Figure 3. Ice Velocity Field Forced by Ocean Current Field of Figure 2.<br />

1041


Ice Transport<br />

Using the current meter data for the winter of 19.76-77 [7] a daily ice velocity<br />

field is calculated for each ice strength. The daily ice motions are then<br />

accumulated to determine ice trajectories over a period of time. To simplify<br />

<strong>com</strong>putations, the dependence of ice velocity at Cape Lisburne on ice strength and<br />

local current velocity is scaled by dimensional analysis. This requires that, in<br />

the ice transport calculations which follow, the ocean current field is assumed<br />

always to vary spatially as shown in Fig. 1. In this current velocity field, the<br />

current speed off Cape Lisburne, v g ' corresponds to the speed at current meter<br />

NC-7. Fig. 4 shows the ice speed, v, calculated from the various daily ice velocity<br />

simulations (e.g., see Fig. 3) at the location of NC-7 as a function of v for ice<br />

strengths p* = 0, 10 4 , and 10 5 N/m. As would be expected, ice velocity g<br />

decreases toward zero with increasing ice strength. It is also observed that for<br />

p* = 10 5 N/m, there is a threshold ocean current of about 0.10 m/s below which the<br />

ice does not move. The threshold is negligible for p* = 104 N/m, and is greater<br />

than 0.5 m/s for p* = 10 6 N/m.<br />

The calculated dependence of ice velocity on current velocity allows us to predict<br />

ice movement through the course of the winter of 1976-77 when the history of ocean<br />

currents is known. From the current meter measurements, we obtain a time history of<br />

v g ' and from velocity field solutions (Fig. 4) for a given value of p*, we<br />

calculate a corresponding time history of ice speeds, v. In determining the ice<br />

velocity, the values of Vg are set to zero if they are northward. Thus, we<br />

prohibit northward ice motion and our results provide an upper bound to southward<br />

0.50<br />

0.40<br />

Iii<br />

E 0.30<br />

:;<br />

"tl<br />

OJ '" Q.<br />

(/)<br />

.l,l '"<br />

1042<br />

0.20<br />

0.10<br />

Ocean Currents Vg Im/s)<br />

Figure 4. Ice Speed as a Function of<br />

Ocean Current at Constant p* .


is 10 5 N/m or greater, but almost all the way to the Bering Strait if the ice<br />

strength is zero. When the strength is 10 4 N/m, ice is transported from Point<br />

Barrow nearly to Cape Lisburne. This analysis prohibits northward ice motions in an<br />

attempt to estimate conservatively whether or not oiled ice can be transported<br />

through the Bering Strait.<br />

When <strong>com</strong>paring our simulated ice motions, we conclude that large southward<br />

transports, such as the one recorded during January 1976 [2], are possible only when<br />

ocean currents are much larger than observed during the winter of 1976-77 and when<br />

the ice along the Alaskan coast is very weak. Therefore, we believe that breakouts<br />

and large-scale southward pack ice motions are a result of a <strong>com</strong>bination of<br />

processes. Deformations must weaken the ice along the coast by the formation of<br />

open water, the ocean currents must reverse their northward flow, and a persistent,<br />

fast (0.4 to 0.5 m/s) southward flow must develop. The process of breakout is more<br />

<strong>com</strong>plex than that envisioned at the inception of this study. It requires that a<br />

hardening/softening plastic model of sea ice be used to account for open water<br />

created during deformations. Because a sequence of events is important for large<br />

motions of the pack ice, simulations should be performed using a time history of<br />

oceanic and atmospheric data to provide the driving forces for pack ice deformation<br />

and transport.<br />

This study represents the first substantial research effort aimed at simulating<br />

Chukchi Sea ice motions on length scales on the order of tens of kilometers with<br />

daily time resolution. Our results <strong>com</strong>plement past studies and provide guidance for<br />

directing future research efforts.<br />

References<br />

1. Pritchard, R. 5., and Reimer, R. W. (1979) "Ice Flow Through Straits," POAC '79,<br />

Vol. 3, Trondheim, Norway, pp. 61-74.<br />

2. Reimer, R. w., Schedvin, J. C., and Pritchard, R. S. (1980) Ice Motion in the<br />

Chukchi Sea, Flow Research Report No. 168, Flow Research Company, Kent,<br />

Washington.<br />

3. Coon, M. D., Maykut, G. A., Pritchard, R. 5., Rothrock, D. A., and Thorndike,<br />

A. S. (1974) "Modeling the Pack Ice as an Elastic-Plastic Material," AIDJEX<br />

Bulletin 24, University of Washington, Seattle, Washington, pp. 1-105-.-----<br />

4. Pritchard, R. S. (1980) "A Simulation of Winter Ice Dynamics in the Beaufort<br />

Sea," in Sea Ice Processes and Models, ed. R. s. Pritchard, University of<br />

Washington Press, Seattle, Washington, pp. 49-61.<br />

5. Coachman, L. K., Aagaard, K., and Tripp, R. B. (1975) Bering Strait, The<br />

Regional Physical Oceanography, University of Washington Press, Seattle,<br />

washington.<br />

1045


6. Tripp, R. B., Coachman, L. K., Aagaard, K., and Schumacher, J. D. (1978) "Low<br />

Frequency Components of Flow in the Bering Strait System," EDS Transactions<br />

Vol. 59, No. 12, American Geophysical Union, pp. 1091.<br />

7. Coachman, L. K., and Aagaard, K. (1981) "Re-Evaluation of Water Transports in<br />

the Vicinity of Bering Strait," in The Eastern Bering Sea Shelf: OceanographY<br />

and Resources, Vol. 1, ed. D. W. Hood and J. A. Calder, u.S. Dept. of Commerce,<br />

National Oceanic and Atmospheric Administration, Office of Marine Pollution<br />

Assessment, Juneau, Alaska, pp. 95-110.<br />

8. Pritchard, R. S. (1981) "Mechanical Behavior of Pack Ice," in Mechanical<br />

Behaviour of Structured Media, ed. A. P. S. Selvadurai, Elsevier, Amsterdam, in<br />

press.<br />

1046


Pierre McComber<br />

NUMERICAL SIMULATION OF ICE ACCRETION<br />

USING THE FINITE ELEMENT METHOD<br />

Universite du Quebec<br />

a Chicoutimi<br />

ABSTRACT<br />

Canada<br />

The rate at which ice grows on a cylinder exposed to supercooled droplets is a func­<br />

tion of the number and sizes of droplets impinging on different points of the surface.<br />

The shape of the ice accretion, however, modifies the air velocities upstream of the<br />

cylinder and therefore the droplet trajectories as well. Because of the resulting<br />

irregular boundary, the finite element technique is appropriate the solve this prob­<br />

lem numerically. To account for the change in shape of the obstacle as the ice grows,<br />

a two-dimensional finite element grid, modified at each step of the time integration,<br />

is used to solve both the air velocity field and the droplet velocity field. From<br />

the droplet speeds, the amount of supercooled water impinging on different points of<br />

the cylinder and the resulting ice thickness are calculated.<br />

Some results of the numerical simulation are <strong>com</strong>pared to samples of ice accretion ob­<br />

tained in a wind tunnel. The <strong>com</strong>parison indicates that the numerical simulation is<br />

adequate to predict the shape and thicknesses of accretions formed in dry growth<br />

conditions.<br />

1. INTRODUCTION<br />

Atmospheric ice accretion on different structures is presently being investigated by<br />

different research groups to solve various environmental problems [lJ. One of these<br />

problems is related to the increasing importance of electrical energy transport lines<br />

in northern regions where they are exposed to adverse weather conditions. In parti­<br />

cular, it has be<strong>com</strong>e important for certain utility <strong>com</strong>panies to gather meteorological<br />

data on this phenomenon. Since the adequate instrumentation is usually not available<br />

where and when it occurs, one has to try to determine the meteorological conditions<br />

of formation from an ice accretion sample. For this purpose, an accurate numerical<br />

1047


+ 0.102 R 0·955<br />

e<br />

0.2 < Re < 2<br />

Equations 3, 4, 5, 6 and 7 are non-linear, and therefore require an iterative method<br />

of solution. A Newton-Raphson sheme was used to obtain convergence of the non-linear<br />

equations. The velocity of air V was taken as the initial velocity for the droplet<br />

a<br />

-+field<br />

V • This corresponds to the case of negligible inertia (K = 0). The calcula-<br />

e<br />

tion then proceeds from lower values of K to larger ones using the results of the<br />

-+-<br />

last iteration as the intitial value of V for the next iteration. With such a mee<br />

thod convergence was obtained in less than four or five iterations for each successi-<br />

ve value of K. This procedure offers the advantage that the results obtained for<br />

increasing values of K can be filed as they are calculated. Since K is the only pa­<br />

rameter involving the droplet diameter in the solution, the results can be readily ex­<br />

tended to a droplet diameter distribution.<br />

Droplets impingement. The droplet velocity field was used to determine local impinge­<br />

ment efficiency S on the cYlinder by the following relation applicable for a «L:<br />

-+- -+-<br />

V .dl<br />

e<br />

for S > 0 (7)<br />

The local impingement efficiency, when multiplied by the liquid water content wand<br />

the free stream velocity Yo' gives the mass flux of water impinging on the surface of<br />

the iced-covered cylinder. The maximum angle of collection 8 m is obtained when S<br />

reaches zero.<br />

Droplet diameter spectrum. Attempts were made at<br />

first to use a statistical distribution to fit the<br />

droplet diameter spectrum. Because of the importan<br />

ce of the few large droplets in the total volume,<br />

for which the statistical distributions tried were<br />

not very accurate, it was better to use directly<br />

the spectrum obtained by experiments as shown in<br />

Table 1. The total water collected by the ice co­<br />

vered cylinder was calculated by a summation of the<br />

collection for different diameters.<br />

Ice accretion volume and shape. Whenever super­<br />

cooled droplets hit an object, they can freeze lo­<br />

cally without runbacks or they can spread on the<br />

surface before freezing. The first condition is<br />

called dry growth and results in the formation of<br />

Mean<br />

diameter<br />

5.1<br />

15.3<br />

25.5<br />

35.7<br />

45.9<br />

56.1<br />

66.3<br />

76.5<br />

86.7<br />

96.9<br />

107.1<br />

117.3<br />

127.5<br />

137.7<br />

TABLE 1<br />

Number of<br />

(Ilm) droplets<br />

945<br />

1093<br />

481<br />

211<br />

74<br />

38<br />

30<br />

13<br />

19<br />

11<br />

10<br />

3<br />

1<br />

1<br />

1051


(a) 0 kV/cm (c) -1 0 kV/cm<br />

(d) -IS kV/cm (e) -20 kV/cm (f) 20 kVrms/cm<br />

Figure 4. - Elongation of air bubbles under applied electric field.<br />

(cutting plane in the cross section of the accretion)<br />

The appearance of large bubbles in ice obtained with a dc negative applied field<br />

(Fig. 3) may be attributed to two factors. The first factor may be the increase of<br />

impact speed of small droplets due to the electrostatic attraction exerted by the<br />

electric field upon the polarization charges of the droplets . This effect should be<br />

ac<strong>com</strong>panied by an increase in the collection efficiency of small droplets. As a re­<br />

sult, the deposit temperature should increase. On the other hand, an increase in the<br />

deposit temperature will produce a decrease in the bubble concentration or an increa­<br />

se in the transmittance [6J; this is inconsistent with the results shown in Fig . 2,<br />

1061


Wilfred R. McLeod,<br />

Tedmical Coosultant<br />

A'lMOOPHERIC SlJPERS'1'ROCTU ICE ACaMUIATICN<br />

MFJ\SURI:MENl'S<br />

Marathon Oil Conpany U.S.A.<br />

Ice acamulatioo neasurements are reported which were rrade 00 Middletoo Island<br />

in the Gulf of Alaska at a 1G-meter elevatioo wring the winter of 1975-1976,<br />

and 00 St. Paul Island in the Bering Sea wring the fall of 1976, the winter of<br />

1976-1977, and the fall and winter of 1978-1979 at 10-, 20-, and 30-meter<br />

elevatioos. These events are reported along with neteorological cooditioos<br />

prevalent before, during and after icing. The data are used to establish<br />

preliminary nethods for design rea:JlllE!ndatioos for offshore a1:!!ospheric icing on<br />

derricks, flare booms, antennas, quarters and superstructures 00 an offshore<br />

platform located in subarctic or Arctic regioos.<br />

1067


This paper attempts to relate the incidence of atmospheric icing to the better<br />

studied sea spray iCing phernneoon. According to Minsk (1977) in his review of<br />

both atmospheric and sea spray icing research, relatively little is known<br />

about atmospheric icing in polar regions, save some continental data which are<br />

inapplicable to the arctic marine environment. Generally, the atmospheric icing<br />

phenomenon is considered less severe than the sea spray icing problem and<br />

hence is ignored. '1t1is paper, through a series of new physical measurements on<br />

two small subarctic islands, seeks to prOlTide a better scientific basis for<br />

determining the physical circumstances that lead to significant atmospheric icing<br />

events. In addition to recognizing those meteorological conditions during which<br />

icing events are likely to occur, we wish also to quantify the rnagnitooe and<br />

frequency of ice aCCl.DllUlation which may be attributed purely to atm:>spheric<br />

icing.<br />

In earlier studies, it often appears that the sea spray and atnDspheric icing<br />

phenanenon are confounded in shipboard measurements. No original work has been<br />

done here in the field of sea spray icing; such data are presented in this<br />

discussion merely to provide a useful reference to the relative rnagnitooe and<br />

conditions for the two distinct types of icing.<br />

CLIMA'IOIffiICAL cnISIDERATICNS<br />

KOnishi and Saito noted that a two-year cycle seemed evident in the wind, ice and<br />

tenperature patterns in the Bering Sea for the period 1960-71. The range of<br />

tenperature variations (based on May values) was on the order of 3.6 0<br />

around 35.6 0<br />

F centered<br />

F, with odd years being warmer and even years cooler. Exceptions<br />

were found in 1965 and 1971. During these periods, the northerly winds also<br />

prevailed as they did in the "low" tenperature even years. Evidently, what is<br />

reflected are differences in the position of the Aleutian low and the North<br />

Pacific high. Generally, storms enter the Bering Sea fran the southwest or<br />

south, sometimes fran Kanchatka on the west, or fran the Gulf of Alaska on the<br />

southeast. Many storms die as they encounter ice fields in winter, others<br />

continue across the Alaska peninsula into the Gulf of Alaska, and sane lIDVe<br />

northeastward up the Kuskokwim and Yukon Rivers (Figure 1).<br />

1068


Sea spray Icing Considerations<br />

we should briefly consider sea spray icing to distin;Juish it clearly fran the<br />

atnospheric icing phenanenon discussed later. At least a portion of the sea<br />

spray icing reported in the literature may, in fact, be due to atnospheric<br />

icing.<br />

Studies of ship icing show that maximum icin;J generally occurs in the rear of low<br />

pressure areas, durin;J north, northwest and west winds; however, a sub-maximum of<br />

icing events may also occur in the forward part of a low with northwest or east<br />

winds. According to Borisenkov and Pchelko (1972), 57% of the 442 reported cases<br />

of ship icing in the Beril'l3 Sea occurred in the rear of a low pressure area,<br />

while 23% of the events were reported in the forward part of the low; all other<br />

cases accounted for 11%. The icil'l3 period defined by these reports is fran<br />

December through March, with corresponding frequency of ship icing of 20%. This<br />

is not so strikil'l3 a low pressure systan dependence as is found elsewhere. In<br />

the sea of Japan, for exanple, 93% of the icing events occur in the rear of a low<br />

pressure systan.<br />

It is even more interesting to note that these icing events reported by Borisen­<br />

kov and Pchelko are almost all clustered about the Pribilofs, with the balance<br />

occurril'l3 to the southeast as far as western Bristol Bay. Only about 11 events<br />

occurred outside this area. Thus, these figures are truly representative of the<br />

study area, the southeast Bering Sea. Certainly these data are biased towards<br />

fishing grounds since the reporting vessels tend to concentrate in these areas.<br />

However, in view of the storm tracks noted and the probable increase in icing<br />

events trailing and leading these storms, it seans likely that the Pribilofs and<br />

the area to the southeast do, in fact, represent more severe icing regions. We<br />

conjecture that a relatively more severe atmosphere icing hazard might be found<br />

here, too.<br />

It sb:luld also be noted that, as can be inferred fran Mertins (1968) data, thick<br />

accumulations of sea spray icil'l3 would most probably result fran strOn;J winds<br />

blowil'l3 for an extended period, 3 - 6 hours, Figure 2. Upper air tE!1peratures of<br />

O· F or less at the 850 mb level are also indicative of atmospheric icing. Thus,<br />

presence of a cold trough at these levels may be taken as a short-term prediction<br />

factor.<br />

1069


location such as the Mscni region (see Table 1, which gives the types of icing<br />

that occurred in three height zones [Glukhov 1972) on a tower at OOOinsk, near<br />

Mscni). On the other ham, the general relationship of rime ice and mixture<br />

aCCl.lDUlation will likely hold true reganUess of location. If one Cbubles the<br />

values measured on the OOOinsk tower at the 25-meter height, a thickness of about<br />

4 an dianeter and a mass of 320 glm result. It is questionable whether the<br />

extreme glaze accumulation of 4 - 6 indies (10 - 15 em) reported in Great Britain<br />

in 1940 would occur offshore since the vertical theD1lal gradient over water is<br />

less extreme than Cl\l'er lam and, therefore, the conditions for superoooled rain<br />

falling en a cold accumulating surface are less likely. Nonetheless, a thickness<br />

of 5 an might reasonably be expected.<br />

Table 1. Occurrence (i) of Type of Ice by Height<br />

SOft rime ice predaninated to a height of 100 m (328 ft).<br />

Type of Ice<br />

Height SOft Glaze Han] Mixture<br />

Range (m) Rime Rime<br />

0-100 23 5 5 2<br />

100-200 29 31 25 23<br />

200-300 48 64 70 75<br />

No. of cases 227 237 633 396<br />

OUr own evaluations of icing rate variation with height are given in Figure<br />

4. The data used the monthly values for the two d:>servation years, 1976-77<br />

and 1978-79. We include in Figure 4 both the linear regression line for total<br />

seasonal accumulation and for mean monthly accumulation. Both of these relation­<br />

ships support the hypothesis of an increase in ice accretion with elevation,<br />

though individual monthly accretions may not reflect this trend. Of course, only<br />

two years of data are insufficient to give anything IIDre than preliminary design<br />

criteria.<br />

CaIplter Analysis of NOM Data<br />

Of several stations considered in our OCIIputer analysis, COld Bay and St. Paul<br />

were chosen as the IIDst typical of the Bering Sea marine environnent. Figure<br />

1071


Coosistently, the dew point depressioo for all events but three shown in Figure<br />

10 lies within 8" F. Are these then legitimate events? In this case, why<br />

does the observer not note icing events at tE!lTperatures aboITe 32" F? What is<br />

evidently needed to verify these data are contact tenperature measurements to<br />

coofirm that either the icing sensor probe is significantly below 32" F, or<br />

continuous air tenperature data to show that the tE!lTperature does not fall<br />

intermittently below freezing in these time intervals notwithstanding the IDurly<br />

data.<br />

Total Precipitatioo Ananalies<br />

Of primary ooncem here is the aRl'lrent discrepancy between the equivalent<br />

precipitatioo recx>rded by the Roseroount detector and the rain gauge 00 St. Paul<br />

(e.g., Figure 8, the events of 29 December 1976, for which ooly a trace is<br />

recx>rded in the rain gauge). 'nle problem of underestimating precipitatioo in the<br />

Arctic is a general ooe. The official precipitation statistics in the canadian<br />

Arctic have already been questioned by Walker and Lake (1975). 'nley have given<br />

values of 9% for the official underestimatioo of rainfall, and 40% for the<br />

corresponding snowfall underestimatioo. Limited Arctic runoff studies also<br />

support such underestimation of precipitatioo by official recx>rds. 'nle cbserver<br />

00 St. Paul states that the dnm-type rain collector is, in all probability,<br />

unusually susceptible to underestimatioo of total precipitation due to the<br />

frequent high winds which blow precipitatioo


Table II<br />

Air Dewpoint<br />

Date Wind '1'EnF. Depression Precipitation<br />

1;7 (3 hrs) E to SE 31· F 2· F SI'rIW<br />

1/11 (6 hrs) E to NE 32· F o· to 1· F snow<br />

1/16 (1 hr) calm 32· F o· snow<br />

1/17 (2 hrs) calm to N 32· F 1· snow<br />

Hence, these criteria, while primitive at this stage, suggest that we could<br />

greatly reduce the number of precipitation events which rust be considered as<br />

possible icing occurrences and concentrate on the ice equivalence problem.<br />

Satellite Ptotographs might also be used to judge an event's ocrurrence with<br />

respect to a given low pressure system's position, as discussed earlier.<br />

'lbe reliability of hindcasting methods for Arctic waters is sanewhat questionable<br />

at this time. Further field measurements in the areas under evaluation where<br />

icing tends to be highly cumulative are desirable in order to verify usable<br />

techniques. Nonetheless, it is possible to attenpt a preliminary approadl to<br />

oatpUter modeling of ice forecasts, Figure 12. Data available include ship<br />

forecasts together with the NOM weather tapes fran the Pribilof Islands, the<br />

Aleutians and the surrounding oontinental statiCl'ls.<br />

Hindcasting of atJrospheric glaze fOITRatiCl'l is sensitive to the accurate pre­<br />

diction of precipitation amount and capture efficiencies of individual structural<br />

members. This requires estimates of precipitatiCl'l particle sizes and knowledge<br />

of CDnditions aloft.<br />

1078<br />

Hindcasting of icing fran fogs similarly requires infoITRation Cl'l the vertical<br />

distribution of liquid water, estimation of the height of low level atmospheric<br />

inversions, and infoITRation giving fog depth versus wind speed.<br />

Hindcasting of icing fran snow accretion will probably need to be based Cl'l<br />

primitive enpirical aSSUllptions about the ability of a structure to build up snow<br />

aCCUlllllation. In general, this may not prove to be a problem as much for diffuse


members (e.g., masts or derricks), as for drifting problems between ooildings or<br />

pieces of dec:k-m:xmted equipnent.<br />

'nle author acknowledges Dr. Arnold Court, Professor of Climatology at california<br />

State University, Northridge, california; and David T. Hodder, Chief Scientist,<br />

Geoscientific Systems and COnsulting, Playa Del Ray, california, for their<br />

assistance in developing the views expressed in this paper. The author also<br />

acknowledges James Pruter and Anna Frisby of the National Weather Service for<br />

their assistance at the Bering Sea Test Site during the data rollection phase.<br />

BIBLIver, NH.<br />

2. Borisenkov, E. P., ed.; I. G. pchelko, ed. (1972) Indicators for forecasting<br />

ship icing (Metodichski ukazaniia po preduprezhdeniiu ugrozy obleden­<br />

eniia sudov), Leningrad, Arkticneskii i antarkticheskii nauchnoissle­<br />

dovatel'skii institut, 81 p. (in Russian).<br />

3. Chaine', P. M. (1972) Estimating the ice accretion hazard, Atnospheric<br />

Environment Service, Toronto.<br />

4. Chaine', P. M. & P. Skeates (1974) Wind and ice loading criteria selection,<br />

Industrial Meteorological - Study III, Toronto<br />

5. Chaine', P. M., A. R. Wayman, and D. A. Bondy, (1975). In Cloud Icing James<br />

Bay and Churchill Falls Power Projects. Industrial Meteorolog ical,<br />

Study VII. Atnospheric Envirorment Service, Toronto.<br />

6. Court, A. (1960). Reliability of hourly precipitation data, Journal of<br />

Geq>hYsical Research, 65 (12), p. 4017.<br />

1079


7. DeAngelis, R. M. (1974) Superstructure icing, Mariners Weather Log, 18 (1),<br />

p. 1-7.<br />

8. Dunbar, M. (1964) Geographical distribution of superstructure icing in the<br />

Northern Hemisphere, Report No. Misc. G-15, Directorate of Physical<br />

Researdl, Defense Research Board, Canada.<br />

9. Glukhov, V. G. (1971) Evaluation of ice loads on high structures from<br />

aerological observations (K otsenka gololednykh nagruzok na vysotnye<br />

sooruzheniia po dannym aerologidleskikh nabliudenni). Trudy, Vol 283<br />

GQJ, p. 3-11, Leningrad, Gidraneteoizdat (in Russian), (Translaticn:<br />

Soviet Hydrology, selected Papers, p. 223-8, Issue No.3, 1971).<br />

10. Guttman, N. B. (1971) Study of worldwide occurrence of fog, thunderstotlDS,<br />

supercooled low clouds and freezing temperatures. NAVAIR 5O-1C-60,<br />

Naval Weather service Coomand.<br />

11. Konishi, R. & M. Saito, (1974) The relationship between ice and weather<br />

conditions in the eastern Bering sea. In: Oceanography of the Bering<br />

Sea, Institute of Marine Science, University of Alaska, Fairbanks,<br />

Alaska.<br />

12. Kuroiwa, Oaisuke (1965) Icing and snow accretion on electric wires, Researdl<br />

Report 123, U.s. ADny Cold Regicns Research and Engineering LaboratoJ:Y,<br />

Hanover, NH.<br />

13. Lenhard, R. W. (1955) An indirect method for estimating the weight of glaze<br />

on wires, Bulletin of the AMS, 36 (1), p. 1-5.<br />

14. McKay, G. A. & H. A. Thoopson (1969) Estimating the hazard of ice accretion<br />

in Canada from climatological data, Jnl. of Applied Met. 8(6), p.<br />

927-935.<br />

15. McLeod, W. R. (1977) Atnospheric Superstructure Ice Accumulation Measure­<br />

1080<br />

ments, Paper No. 2950, Offshore Tedlnology Conference, Houstcn, Texas.


16. Mertins, H. o. (1968) Icing on fishing vessels due to spray, Marine Observer<br />

32(221), p. 128-30.<br />

17. Minsk, L. D. (1977) Ice accumulation on ocean structures, CRREL Report<br />

77-17 , U • S • Army Cold RegiO'1s Researcn and Engineering Laboratory ,<br />

Hanover, NH.<br />

18. Walker, E. R. & R. A. Lake (1975) Rmoff in the canadian Arctic Ardlipelago.<br />

In: Climate of the Arctic, University of Alaska, Fairbanks, Alaska.<br />

1081


9<br />

INPUT WEATHER DATA FROM STATIONS SURROUNDING THE STUDY<br />

6-HOURLY WINDS<br />

AIR AND DEWPOINT TEMPERATURES (36°ro 20" F)<br />

VISIBILITIES AND CEILINGS<br />

PRECIPITATION AMONTS AND TYPES<br />

MEAN MONTHLY SURFACE WATER TEMPERATURES (NOAA)<br />

8 IF SPRAY, IS TEMPERATURE<br />

IN GLAZE<br />

OR RIME FORMATION RANGE?<br />

COMPILE ICE<br />

ACCRETION RATES<br />

APPLY CORRECTION FACTOR FOR ELEVATION,<br />

ICING THICKNESS, WIND SPEED,AND<br />

STRUCTURAL MEMBER TYPE<br />

PROJECTED ICING HINDCASTING PROCEDURE<br />

Figure 12<br />

1093


I. K. Hill, Head of Hydraulic<br />

Department<br />

A. B. Cammaert, Project Engineer<br />

D. R. Miller, Naval Architect<br />

ABSTRACT<br />

A LABORATORY STUDY OF HEAT TRANSFER<br />

TO AN ICE COVER FROM A WARM WATER DISCHARGE<br />

Acres Consulting Services<br />

Acres Consulting Services<br />

Arctic Pilot Project<br />

Canada<br />

Canada<br />

Canada<br />

A warm water discharge has been proposed to limit ice buildup around an Arctic<br />

liquid natural gas terminal. One approach consists of high velocity jet diffusers<br />

inside a floating curtain surrounding the termlnal area, resulting in the suppression<br />

of ice growth.<br />

Water temperatures and velocities were predicted from a thermal plume model.<br />

Calculation of the effect of the flow on ice thickness requires knowledge of the<br />

rate of heat transfer between the water and ice. Experimental flume observations<br />

were made to confirm the relatlonship used for the heat transfer analyses.<br />

This paper describes the experimental program conducted in an ice flume at the Acres<br />

laboratories. The heat transfer rate was obtalned by measuring the temperature<br />

profile through the ice cover and the change in ice thickness with time.<br />

Tests were conducted with both freshwater and saline ice. The parameters of ice<br />

thickness, water depth, water velocity and water temperature were varied. The test<br />

data are presented and <strong>com</strong>pared with the analytical results.<br />

INTRODUCTION<br />

The Canadian High Arctic is anticipated to contain significant reserves of gas and<br />

oil. In late 1976 the Arctic Pilot Project (APP) was formed, bringing together<br />

Petro-Canada, Nova - an Alberta Corporation, Melville Shipping, and Dome Petroleum.<br />

1094


The concept of the Project was a gas transportation system on the smallest scale<br />

possible which would be <strong>com</strong>mercially viable. It is intended to evolve and<br />

demonstrate appropriate technology and to provide definitive cost and performance<br />

information for facilities in the high Arctic.<br />

The system throughput will be 7 x 10 6 m 3 gas per day, obtained from the Drake Point<br />

field on the Sabine Peninsula of Melville Island.<br />

The Project requires two liquid natural gas vessels designed and constructed to the<br />

requirements of Arctic Ice Class 7 vessels. At intervals of 9 to 16 days, depending<br />

on ice conditions, these vessels will transport the liquefied natural gas by way of<br />

Parry Channel and Baffin Bay to a receiving terminal on the east coast.<br />

At the time of writing, the Arctic pilot Project has made an application to the<br />

Canad1an Government for permission to export the natural gas. An environmental<br />

hearing was held in Resolute, Northwest Territories, in April 1980 to consider the<br />

environmental impact north of 60 0 N. This report has been received and has concluded<br />

that the impacts are acceptable provided certain conditions are <strong>com</strong>plied with by the<br />

Project proponents. The projected start-up date of the Project is early 1986.<br />

GENERAL LOCATION AND ENVIRONMENT<br />

Melville Island is located in the Parry Island group of the Queen Elizabeth Islands.<br />

On the southeastern portion of the island at 75 0 N, 108 0 50'W, is Bridport Inlet, the<br />

site of the terminal facilities. The average mean temperature in February is about<br />

_35 0 C, and in July about +4 0 C, with winter temperatures going as low as _52 0 C<br />

for up to 3 days at a time. A typical year at Bridport Inlet would have about<br />

6,440 freezing degree C days.<br />

Ice problems will be very severe for year-round operation of a marine term1nal at<br />

such a northern location. with the exposure of open water after each ship passage,<br />

ice production is accelerated resulting in a greater volume of ice than under un­<br />

disturbed conditions. Uneven distribution of ice can cause manoeuvering difficulty<br />

if the vessel attempts to berth unassisted.<br />

A previous publication (Carnrnaert et aI, 1979) has ident1fied and quantified this<br />

augmented growth. An active ice management system was required to control or in­<br />

hibit the ice growth to some reduced thickness, allowing the vessels to dock<br />

successfully. It was concluded that the most promising method for this<br />

1095


The layer of moving water was expected to be a few metres thick and held under the<br />

ice by a very weak buoyancy force induced by additional heat added by the jets.<br />

Extensive literature is available on heat transfer to a flat plate (L.C. Thomas,<br />

1980) under various boundary conditions. The behavior of an ice-water interface is<br />

modified ,however ,by the melting or freezing process which,in addition to mass<br />

transfer,results in local changes to the density of the fluid at the interface.<br />

Literature available on heat transfer to ice largely relates to fully developed<br />

turbulent flow such as a river (Cowley and Lavender, 1974).<br />

A <strong>com</strong>mon assumption in heat transfer analyses is that the coefficient of thermal<br />

expansion is independent of temperature in the boundary layer. This is not valid,<br />

for a water-ice interface at salinities of around 25 0/00. At this salinity the<br />

coefficient of thermal expansion at the freezing point is zero and is negative at<br />

still lower salinities.<br />

In the prototype a temperature differential between the bulk flow and the interface<br />

of the order of 2 0 C is expected and the depth of the moving layer is about 3 m<br />

for typical jet designs. In the first 100 m of the ice management zone the under­<br />

ice boundary layer will thicken to between 1 and 2 metres.<br />

The model tests were designed to determine whether the standard heat transfer<br />

formulae developed for other situations and used for the feasibility analyses of<br />

this Project (Acres, December 1978) were sufficiently reliable to ensure basic<br />

feas1bility of the Project. It was recognized that to obtain data for design<br />

purposes, additional test1ng beyond the work described might well be required.<br />

For forced convection with buoyancy effects the system can be described by the<br />

geometric shape, Reynolds number Re, Prandtl number,Pr,and the Grashof number,Gr<br />

which relate viscous to buoyancy forces. The heat transfer coefficient is<br />

described in terms of the Nusselt number,Nu. In the test program water was used<br />

so that the Prandtl number,which is dependent on fluid properties only,was similar<br />

in the model and prototype. In the prototype the longitudinal Reynolds number<br />

ranges of order 10 7 , implying a fully turbulent boundary layer over most of the<br />

area,with an area close to the dock of potential laminar flow, particularly if the<br />

final design used low velocities. Dynamic and thermal similarity require a laboratory<br />

Reynolds number in the same range, that is, either turbulent or laminar. Because<br />

of the initial turbulence in the water the transition Reynolds number in the flume<br />

will be close to the lower limit of 3 x 105. Tests were conducted either side of<br />

this transition value.<br />

1099


The relative importance of buoyancy vis-a-vis forced convection is similar for<br />

similar values of the dimensionless ratio Gr/Re 2 which for the same fluid reduces<br />

to similarity of 6T.L/V2 where L and V are characteristic lengths and velocities<br />

and 6T is the temperature differential. This is equivalent to similarity of the<br />

densiometric Froude number or the Richardson number. Because of the nonlinear<br />

temperature density relationship in water it was felt desirable to maintain similar<br />

temperatures in the model and prototype. Similarity of the ratio Gr/Re 2 then<br />

required that the velocity scale be reduced as the square root of the length scale.<br />

This implied, if large values of Re were to be maintained, that as large a test<br />

facility as possible would be required. Because of the high cost of meeting this<br />

requirement in full, tests were carried out over the desired ranges of Re and<br />

Gr/Re 2 separately without covering the extremes of high Re and high Gr/Re 2<br />

simultaneously. Because of the nonlinearity of the density temperature relationship<br />

6T was varied independently. Tests were also conducted using freshwater and salt­<br />

water, which results in a large variation in Grashof number as indicated by a change<br />

from destabilization of the flow by cooling for seawater to a stabilizing of the<br />

flow in freshwater. The intent of the range of tests selected was to determine<br />

whether,for the conditions expected in the prototype,the preliminary analyses used<br />

were adequate for feasibility. The actual range of parameters examined is given in<br />

Table 2.<br />

Parameters<br />

Velocity (m/s)<br />

Ice Thickness (cm)<br />

Water Temperature<br />

(oC)<br />

Ice Growth Rate<br />

(crn/h)<br />

Temperature Gragient<br />

Through Ice ( C/m)<br />

Nusselt number<br />

Reynold's number<br />

Heat Transfer<br />

Coefficient,<br />

W/m2 oc<br />

TABLE 2 - RANGE OF TEST PARAMETERS EXAMINED<br />

Smooth Freshwater<br />

0.214 to 0.016<br />

17.8 to 3.3<br />

5.20 to 0.36<br />

0.410 to -1.811<br />

211 to 68<br />

6,688 to 411<br />

633,700 to<br />

24,400<br />

988 to 47<br />

Rough Freshwater<br />

0.269 to 0.116<br />

18.0 to 6.9<br />

1.96 to 0.20<br />

-0.119 to -1.500<br />

74 to 36<br />

10,590 to 3,096<br />

764,500 to<br />

325,700<br />

1,200 to 350<br />

Sallne<br />

0.225 to 0.020<br />

20.5 to 6.1<br />

4.79 to 1. 25<br />

-0.643 to -4.500<br />

149 to 5<br />

6,483 to 1,009<br />

675,800 to<br />

64,900<br />

706 to 112<br />

The test runs were carried out by initially forming a solid cover to the desired<br />

thickness with three strings of temperature probes frozen into the cover as<br />

indicated in Figure 1.<br />

1100


Haggkvist, Kenneth<br />

Research Engineer<br />

ABSTRACT<br />

COMBINATION OF A SINKING WARM WATER DISCHARGE AND<br />

AIR BUBBLE CURTAINS FOR ICE REDUCING PURPOSES<br />

Water Resources Engineering<br />

University of Lulea<br />

WREL<br />

Sweden<br />

The work described in this report indicates that a considerable ice reduction effect<br />

in a limited area can be obtained by <strong>com</strong>bining a sinking warm water discharge and<br />

a bottom located air bubble discharge. A laboratory experiment and a theoretical<br />

analysis are presented. In an example based on the conditions in Lulea Harbour<br />

in northern Sweden it is shown that over an area of 50 by 500 metres, the ice<br />

thickness can be reduced to less than 0.2 metres, which is less than 25 % of the<br />

maximum natural ice thickness.<br />

INTRODUCTION<br />

Harbours in northern Sweden (northern Bothnian Gulf) are since 1970 open for yearround<br />

navigation. However, due to repeated breaking and freezing of the ice masses<br />

in fairways and at quays, navigation problems occur during the winter period. The<br />

problems are accentuated in quay areas, where berthing manoeuvres are made difficult<br />

and time-consuming.<br />

Two methods, frequently used for ice reducing purposes in limited areas, are respectively<br />

surface discharge of warm water and transport of "warmer" bottom water to the<br />

surface by means of a bottom located air discharge.<br />

The harbours of the northern Bothnian Gulf are mostly situated in shallow river estuaries,<br />

with very low salinity and homogenous water temperature very close to the<br />

freezing point from surface to bottom. The methods mentioned above are separately not<br />

effective for ice suppression. The surface discharge of warm water is rapidly mixed<br />

with the surrounding water, thereby being heavier than the ambient water, and sinks<br />

to the bottom. Further, the "warm" bottom water, necessary for an air discharge<br />

arrangement to be effective, is not available. If, on the other hand, the two mentioned<br />

methods are <strong>com</strong>bined, a portion of the released sinking warm water can be brought to<br />

the surface by means of a (bottom) air discharge.<br />

In order to make a study of the <strong>com</strong>bination of a sinking surface water discharge and<br />

air-bubble curtains a laboratory experiment has been performed at WREL (Division of<br />

Water Resources Engineering, University of Lulea).<br />

1104


a) Run 1; Two parallell air curtains<br />

B=O.5mi [):O.5mi 8/0=1<br />

A-A<br />

b) Run 2; One air bubble curtain parallell to the flume wall.<br />

Air pipes<br />

... Surface discharge of water heavier than ambient (flume) water.<br />

=i Ambient (flume) flow.<br />

x Downstream distance from discharge<br />

B Width between air-pipes<br />

D Flume water depth<br />

The water volume between the air-pipes and the air-pipe/flume wall respectively is<br />

below termed the box.<br />

Before (and after) each run the ambient water temperature was measured with operating<br />

air curtains. The discharge of colder water was then started and its temperature was<br />

measured. When steady state conditions were reached the water temperature was measured<br />

over the flume width at three different depths and for various downstream distances.<br />

Listed below are the numerical values of the parameters characterizing the experiment.<br />

Flume water depth<br />

Flume water velocity<br />

Sinking plume discharge rate<br />

Initial temperature difference between discharge<br />

and ambient water<br />

Initial density difference<br />

1106<br />

0.5 m<br />

0.01 - 0.02 m/s<br />

0.3.10- 3 m 3 /s


G.D. Fonstad<br />

R. Gerard<br />

B. Stimpson<br />

THE EXPLOSIVE DEMOLITION OF<br />

FLOATING ICE SHEETS<br />

Hydraulic Engineer, River Engineering Branch<br />

Alberta Environment<br />

Professor, Dept. of Civil Engineering<br />

Associate Professor, Dept. of Mineral Engineering<br />

University of Alberta<br />

Edmonton<br />

Alberta<br />

Canada<br />

Abstract<br />

Demolition of floating ice sheets is a <strong>com</strong>mon technique used to clear shipping<br />

lanes, construct temporary port facilities in Arctic and Antarctic environments and<br />

to mitigate ice jam effects on inland waterways both before and after ice jam formation.<br />

Mellor carried out a review and analysis, on the data existing to 1972, of the<br />

effects of point charges on floating ice sheets. On the basis of this analysis Mellor<br />

made preliminary re<strong>com</strong>mendations of the optimum charge size and placement depth as a<br />

function of ice thickness.<br />

In this paper a series of tests conducted to confirm Mellor's analysis and to<br />

determine the optimum spacing of charges in a row are described. The appropriate dimensionless<br />

terms are derived, and equations giving the optimum ice sheet demolition<br />

parameters are given.<br />

Introduction<br />

Ice has an adverse impact on the operation of ports and navigable waterways.<br />

Various methods of keeping ports open have been developed over the years. These include<br />

icebreakers, explosive destruction of the ice and, more recently, the use of<br />

air cushion vehicles. In Arctic and Antarctic environments explosives have been used<br />

in the clearing of channels and the construction of temporary port facilities.<br />

On inland waterways in cold regions ice jams are a constant concern. In many<br />

instances efforts are made to weaken the ice cover at critical locations by blasting<br />

in advance of breakup. If an ice jam does form, blasting is often helpful in removing<br />

the jam.<br />

In carrying out such explosive demolitions it is desirable to have some knowl-<br />

1114


and can be incorporated into the constants implicit in a function such as Equation 2.<br />

It has been shown [3] that the energy released upon detonation is proportional to the<br />

weight W of the explosive charge, and can thus be substituted for the energy E. Also,<br />

experience has shown that, except perhaps for very shallow depths, the depth of water<br />

has little influence on the results.<br />

Hence Equation 2 can be reduced to:<br />

R _ (ti' d , 1 ,do, Ro)<br />

K2 - f K2 K2 gK 2 K2 K2<br />

••• (3)<br />

where K2 = w 1 / J , which is the well known 'cube root' scaling criteria for blast effects,<br />

applied to the case of floating ice.<br />

Information Available in the Literature<br />

Although explosives have been used for clearing ice for over 200 years [4],<br />

little information could be found which dated prior to the 1960's. Cole [1] laid the<br />

groundwork for ice demolition investigations through his study of underwater explosions,<br />

but it was not until the 1960's that reported experimentation on ice demolition<br />

was conducted in the western world. During the 1960's and early 1970's a number<br />

of such studies were reported. Those prior to 1968 have been reviewed by Bolsenga [4].<br />

Other information was included in the data <strong>com</strong>pilation by Mellor [5], and some additional<br />

results are given by Nikolayev [6].<br />

Four cratering mechanisms can be identified from this past work. First, the<br />

impact of the shock wave may cause <strong>com</strong>pressive failure of the ice sheet [6], though<br />

that this occurs is not generally agreed upon. For instance, Barash [7] notes that<br />

the shock wave causes the ice to "rise in the shape of a dome". Kurtz et.al. [2]<br />

noted a similar ice dome but from analysis of the high speed photography of their experiments,<br />

considered that the ice may not be shattered by the passage of the shock<br />

wave, as there was no indication of the upper ice surface spa1ling. Second, the gas<br />

bubble vents through the ice sheet to the atmosphere. At this time, ice is thrown out<br />

either "mostly upward or mostly radially depending on the phase of the gas bubble"[7].<br />

The third cratering mechanism is a water surge through the crater following the venting<br />

gases. As the gas bubble vents, water rushes in to fill the void, and fluid momentum<br />

causes it to surge through the crater. A fourth and final mechanism has been noted<br />

though only for 'deep' charge placements in shallow water: this is a second water<br />

surge through the ice which is "always very muddy" [2].<br />

It has been noted in the literature that even with these several means of ejecting<br />

ice from the crater, upwards of 90% of the ice still remains in it.<br />

These early studies determined by trial the charge weight and the optimum depth<br />

1116


on Drummond Lake in the Chilcotin region of central British Columbia, on ice which<br />

was approximately 0.37 m thick. In all some 86 separate single shots were fired.<br />

These included four explosive types: a military plastic explosive with a PETN base,<br />

for the majority of the tests; and three <strong>com</strong>mercial explosives manufactured by<br />

Canadian Industries Limited.: 40% Forcite, Amex II and Hydromex, which are respectively<br />

a dynamite, a blasting agent and a bulk slurry explosive.<br />

Analysis<br />

In order to analyse how well Mellor's regression equation predicted the independent<br />

Chilcotin data it was considered insufficient to scale from the curves given<br />

by Mellor, especially with his re<strong>com</strong>mendation for caution. The data used by Mellor<br />

was collected and re-analysed by regression using the same model as Mellor. During<br />

this repeat analysis it was confirmed that the terms with coefficients b 1 and b 4<br />

could indeed be omitted as their contribution to the multiple correlation coefficient<br />

was minimal. The analysis predicted a mean scaled crater radius of 2.1? m/kg 1 / 3 , with<br />

a standard error of estimate of 0.55 m/kg 1 / 3 , and a multiple correlation coefficient<br />

of 0.69, which is <strong>com</strong>patab1e to that found by Mellor, as it should be.<br />

The regression equation obtained was:<br />

2 2 3 2<br />

Y = 1.55 + 7.73{X 2 ) - 0.419{X 1 ) - 14.8{X 2 ) + 0.141{X 1 ) - 0.551{X 1 X 2 )<br />

2 3<br />

+ 1.32{X 1 X 2 ) + 6.59{X 2 ) (6)<br />

This equation was used to calculate scaled crater radius for Mellor's data, for<br />

which the <strong>com</strong>parison between predicted and observed crater radius is shown in Figure<br />

2. The equation underpredicts larger values of scaled crater radius and overpredicts<br />

the smaller ones. The skew of the data about the expected best fit slope of. 1.0 indicates<br />

that there may be a factor unaccounted for in the analysis. An attempt was<br />

made to take this skew out by including the gravity term of equation 3 in a separate<br />

regression analysis. There was, however, only a slight improvement in the correlation<br />

coefficient, and the skew remained. The Chi1cotin data were analysed in similar<br />

fashion, with the results shown in Figure 3. The three lowest Hydromex shots were<br />

omitted from the analysis for the line of best fit, as three other shots out of eight<br />

total Hydromex shots had misfired, and it is considered that the three lowest shown<br />

in Figure 3 had only partially detonated. From Figure 3 it can be seen that Equation<br />

6 generally - overpredicts - the scaled radii observed and there is a suggestion that<br />

the same skew noted in Figure 2 exists. At the time of writing, attempts to explain<br />

the general overprediction and to improve the relationship are continuing.<br />

1119


for both the Hydromex and the Amex misfired, even though careful attention was paid<br />

to ensuring the water integrity of the charges. It is thought that either water seepage<br />

into the charges, or insufficient loading density caused these charges to misfire.<br />

Depending on the operation undertaken, though blasting agents and slurries are less<br />

expensive than other explosives, the care required to ensure the charges are watertight<br />

may detract from the speed at which the operation can be conducted.<br />

For any application of the optimum ice demolition conditions given herein, it<br />

is re<strong>com</strong>mended that a few trial shots be made before undertaking the main program.<br />

Acknowledgements<br />

The writers would like to express their appreciation to the Canadian Defence<br />

Research Establishment - Suffied, who financed and assisted in the investigation. A<br />

special thanks is due to Mr. G.K. Briosi, of that establishment, who assisted in the<br />

data collection throughout the field experiments. The field tests were carried out by<br />

1 Combat Engineer Regiment, Canadian Armed Forces, under the technical supervision of<br />

the writers. Mrs. S. Notton, of Alberta Environment drafted the figures.<br />

References Cited<br />

1. Cole, R.H.,(1948),'Underwater Explosions', Princeton University Press.<br />

2. Kurtz, M.K., R.H. Benfer, W.G. Christopher, G.E. Frankenstein, G. Van Wyhe and<br />

E.A. Roguski ,(1966), 'Consolidated Report, Operation Breakup,<br />

FY-66, Ice Cratering Experiments, Blaire lake Alaska', NCG/TM<br />

66-7, U.S. Army Nuclear Cratering Group, lawrence Radiation<br />

laboratory, livermore, California.<br />

3. Baker, W.E., P.S. Westine and F.T. Dodge,(1973),'Similarity Methods in Engineering<br />

Dynamics', Spartan Books, Hayden Book Co. Inc., Rochelle Park,<br />

New Jersey.<br />

4. Bolsenga, S.J.,(1968),'River Ice Jams', Research Report 5-5, U.S. Army Corps of<br />

Engineers, lake Survey District, Detroit.<br />

5. Mellor, M.,(1972),'Data for Ice Blasting', CRREl Technical Note, U.S. Army Corps<br />

of Engineers, Cold Regions Research and Engineering laboratory,<br />

(CRREl), Hanover, New Hampshire.<br />

6. Nikolayev, S.Ye.,(1970),'Blasting Fast Ice in the Antarctic', Soviet Antarctic<br />

Expedition, Information Bulletin. (Translation by the U.S. Army<br />

CRREl, February, 1973).<br />

1122


7. Barash, R.M., (1966),'Ice Breaking By Explosives', NOLTR 66-229, U.S. Naval<br />

Ordnance Laboratory, White Oak, Maryland. (Extracts from a<br />

Confidential report, 1962,'Underwater Explosions Beneath Ice',<br />

NOLTR 62-96, U.S. Naval Ordnance Laboratory, White Oak, Maryland.<br />

8 Personal Communication,(1980), Dr. M. Mellor.<br />

1123


D. B. Coveney<br />

Research Officer<br />

ABSTRACT<br />

CUTTING ICE WITH<br />

"HIGH" PRESSURE<br />

WATER JETS<br />

National Research Council<br />

of Canada<br />

Canada<br />

A high pressure jet of water can be used to cut a slot into or through<br />

a sheet of ice, thereby substantially weakening the ice sheet. Such weakening<br />

could be particularly useful to enhance the ice breaking capabilities and/or to<br />

reduce the overall power and fuel requirements of an ice breaking vessel. Although<br />

water jet cutting is less efficient in material removal than mechanical modes of<br />

cutting, its ability to cut with a substantial mechanical stand-off from the ice<br />

sheet and with a concentrated, high level of power input into the ice would provide<br />

significant practical advantages for the water jet cutting method.<br />

This paper describes the ice cutting performance of small to moderate<br />

scale water jets in fresh water ice and of small scale water jets in a simulated<br />

sea ice. The majority of cuts produced a narrow, clean kerf, indicative of erosion<br />

in a ductile material, while other cuts produced a wide spalled trench, indicative<br />

of spalling in a brittle material. Still others produced a <strong>com</strong>bination of the two<br />

modes of cutting, with a wide, shallow trench and a narrow, deep kerf below the<br />

trench. The causes and the effects of these characteristics on ice cutting per­<br />

formance are discussed, along with the effects of jet traverse speed, nozzle<br />

diameter, nozzle pressure, nozzle stand-off, ice characteristics and the overall<br />

scale of the system. An empirical relationship, derived by regression analysis, is<br />

presented correlating the jet penetration to the power in the jet, the jet traverse<br />

speed, the nozzle stand-off and the estimated ice temperature.<br />

1124


1.0 INTRODUCTION<br />

Cutting a slot into a sheet of ice can reduce its flexural strength con­<br />

siderably. Such a slot or multiple slots should be useful in easing the passage of<br />

an ice-breaking vessel through ice fields. The substantial weakening of an ice<br />

sheet by cutting one or more grooves in the ice by means of a high pressure water<br />

jet has been proposed as a possible means of extending current ice breaking capabil­<br />

ities and reducing fuel consumption. A relatively simple device, the high pressure<br />

water jet, used as a cutting tool, has the potential for development into a rugged,<br />

practical system for notching ice ahead of an ice-breaking vessel. Although mech­<br />

anical modes of cutting can remove material more efficiently, a water jet has the<br />

advantage of non-mechanical contact and can cut with a substantial stand-off from<br />

the material being cut. This characteristic along with the ability to introduce a<br />

concentrated, high level of power into the material would provide significant prac­<br />

tical advantages for the water jet cutting method when used to assist ice breaking.<br />

Previous work by the Gas Dynamics Laboratory of the Division of Mechanical<br />

Engineering of the National Research Council of Canada in cutting a variety of mater­<br />

ials with high pressure water jets and a few water jet cuts in ice at the University<br />

of Missouri at Rolla during frozen soil cutting trials for the U. S. Army Cold<br />

Regions Research and Engineering Laboratory [1] led to exploratory small scale ice<br />

cutting trials in the Gas Dynamics Laboratory [1]. While these initial trials<br />

showed that ice indeed could be cut with high pressure water jets, extrapolation of<br />

the results to a full scale system was impractical. After a subsequent series of<br />

field tests by CRREL at very high pressures [1], the Gas Dynamics Laboratory in col­<br />

laboration with CRREL made various series of cuts in ice ranging from floating<br />

ice [2] to manufactured ice to lock wall ice collars with a pumping system about one<br />

full order of magnitude larger than the laboratory system. These cuts covered a<br />

fairly wide range of conditions, from relatively high speed shallow penetration cuts<br />

to low speed relatively deep penetration cuts. Extrapolating about two orders of<br />

magnitude from these results, while not at all reliable, did indicate that a realis­<br />

tic full scale system might be possible.<br />

To further investigate the potential cutting ability of water jets in ice,<br />

larger scale field tests were initiated [3] [4][7] by the Low Temperature Laboratory<br />

of the Division of Mechanical Engineering of the National Research Council of Canada<br />

and conducted in collaboration with the Gas Dynamics Laboratory. Through the course<br />

of this investigation, the field tests were supplemented by further fairly small<br />

scale laboratory tests [5][6] including one series of cuts in a simulated sea<br />

ice [6].<br />

1125


2.0 ICE CUTTING SYSTEM<br />

Cutting ice with a water jet is achieved by impacting a high velocity jet<br />

of water onto the ice. The resulting velocity and directional changes apply forces<br />

to the ice sufficient to fracture some of the weaker bonds between and within crys­<br />

tals. By traversing the jet across the surface of the ice a slot can be cut into<br />

or even through the ice. Figure 1 shows a water jet cutting such a slot in a<br />

floating ice sheet, and Figure 2 shows the kerf cut by the water jet.<br />

1126<br />

Figure 1: Water Jet Cutting of<br />

Floating Ice Sheet<br />

Figure 2: Kerf Cut by<br />

Water Jet<br />

Typically, the jet is produced by accelerating high pressure water<br />

through a convergent steel nozzle. The high pressure water is supplied by a pump­<br />

ing system usually drawing the water for the jet from under the ice sheet . For<br />

most of our field testing the swing of a hydraulic crane with a telescoping boom<br />

was used to traverse the cutting nozzle.<br />

Suitable and accessible test sites were selected, on a spring-fed pond<br />

for the first series of field t ests (March 1977) [3] and on the Ottawa River for<br />

the second and third series (February 1978 [4] and February 1979 [7] respectively).<br />

For the field tests the ice thickness ranged up to about 0.7 metre.


frequently thrown considerable distances; the small particles usually were ejected<br />

in the spray of spent water from the jet. Except for the shallowest cut all cuts of<br />

the third series resulted in a narrow fairly clean kerf about 13 mm wide, similar to<br />

the cuts of the first series.<br />

Of the measurable cuts in the ice blocks, most produced a spalled groove,<br />

a few produced a recognizable kerf and the remainder simply melted a shallow groove<br />

in the ice. It was noticed that kerf cuts only occurred at the higher nozzle pres­<br />

sures (above 48 MFa) and at the higher ice temperatures (above -SoC), while spalling<br />

occurred between 7 and S2 MPa nozzle pressure and a melted groove was often found at<br />

nozzle pressures from 4 to 21 MFa.<br />

Cuts in the laboratory fresh water ice sheets varied from deep, narrow,<br />

clean kerf cuts at the higher nozzle pressures to shallow, widely spalled cuts at<br />

the lower nozzle pressures. No sharply defined change in mode occurred; rather,<br />

there was a gradual increase in spalling as the pressure was lowered.<br />

For the simulated sea ice most of the cuts were essentially clean kerf<br />

cuts with a small degree of surface spalling. However, for the first seven tests,<br />

as the pressure was reduced below about 40 MFa, the spalling became wider and<br />

deeper until at about 13 MFa only a spalled trench was produced.<br />

4.0 ICE CHARACTERISTICS<br />

The fresh water ices cut in the field tests were generally a type of nat­<br />

ural ice <strong>com</strong>monly found on lakes and rivers, having many layers of snow ice with an<br />

underlying layer of clear ice. For the first series the ice temperature was OoC<br />

and the top layers had candled, while for the second series the ice near the top<br />

surface varied from about -lloC to _2 o C and for the third series it ranged from<br />

about -18 o C to _lloC. The colder ice was harder and stronger and obviously more<br />

difficult to cut, while the candled top layers of ice cut eaSily.<br />

All the fresh water ice made in the laboratory was clear with the charac­<br />

teristics of ice grown unidirectionally from the free surface. The blocks of ice,<br />

having been stored in a chest freezer, were at a uniform temperature throughout<br />

varying from _18 o C to OOC. However, for the July/August 1978 tests, the three sep­<br />

arate ice sheets, near their top surface, ranged from about _17 o C to OOC.<br />

There was a variety of ices cut in the tests from the early studies of<br />

NRC and USA CRREL [1][2], ranging from ice blocks to floating ice sheets to lock<br />

wall ice collars. However, all were apparently at or very near to OOC and there­<br />

fore relatively easy to cut.<br />

The saline ice produced for these tests was a first approximation facsim­<br />

ile of first year sea ice. Its salinity at 5 ppt was in the same range as the<br />

"typical average figure" of 4 ppt cited by Pounder [8]. With its many inclusions<br />

1128


and underlying fragile structure, this ice should have been more susceptible to<br />

water jet cutting than fresh water ice.<br />

5.0 ANALYSIS OF TEST DATA<br />

For the cuts in fresh water ice and for those in simulated sea ice, separ­<br />

ate relationships between the jet penetration and the jet parameters have been<br />

derived by multiple linear regression analyses.<br />

From the general expression<br />

Y = f (u, d, p, s)<br />

where: Y average penetration (cm)<br />

u = nozzle traverse speed (km/h)<br />

d nozzle diameter (mm)<br />

p nozzle pressure (MFa)<br />

s = average nozzle stand-off distance (cm)<br />

and assuming that a zero value for this function would result in a zero depth of<br />

cut, a first approximation was obtained by applying multiple linear regression anal­<br />

ysis to a logarithmic transformation of this expression to yield ultimately an<br />

equation of the form:<br />

(1)<br />

E<br />

s (2)<br />

For the cutting of fresh water ice, analyses of this type were conducted<br />

both on individual series of tests and on <strong>com</strong>binations of data from groups of test<br />

series, including published and unpublished data from the early studies of NRC and<br />

USA CRREL [1] [2]. These analyses yielded exponents for nozzle traverse speed con­<br />

sistently near -0.5. Whenever the data covered a sufficient range of nozzle diam­<br />

eter and pressure, the exponent for nozzle diameter tended to range between 1.5 and<br />

2, while that for nozzle pressure tended to vary about 1.5. There was a general<br />

lack of useful correlation to nozzle stand-off.<br />

It was recognized that the exponents for nozzle diameter and pressure were<br />

close to those that appear in the relationship describing the physical jet property<br />

of hydraulic power,<br />

HP = C d 2 p3/2<br />

where: HP hydraulic power (kW)<br />

C dimensional constant<br />

d nozzle diameter<br />

p nozzle pressure<br />

With the simple relationship,<br />

ru<br />

Y = f (HP)<br />

(mm)<br />

(MFa)<br />

consistently good, highly significallt correlations were obtained both for individual<br />

(3)<br />

(4)<br />

1129


clear demarcation between kerfing and spalling, the results of this test program<br />

suggest that about 40 MPa was needed to cut a kerf without excessive spalling in<br />

either fresh water ice or in the simulated sea ice. Still higher pressures gener­<br />

ally produced cleaner cuts. For equivalent conditions a spalled cut tended to be<br />

shallower than a kerf cut. A few small scale cuts simply melted a groove in the<br />

ice.<br />

For the first few simulated sea ice tests, when the ice was still cold,<br />

cutting through the hard surface layer, in all cases, resulted in some degree of<br />

surface spalling, be<strong>com</strong>ing more pronounced as the nozzle pressure was reduced until,<br />

at 13 MPa, only spalling occurred. Cutting into or through the underlying softer<br />

ice produced clean kerf cuts.<br />

For the fresh water ice cutting regression analysis, equation (5) provides<br />

a statistically excellent fit to the entire body of data. A <strong>com</strong>parison of the slope<br />

of equation (6) to that of equation (5) indicates that penetration in the simulated<br />

sea ice was more than double that in fresh water ice when the jet parameters were<br />

similar. Apparently, the brine and air inclusions within the saline ice structure<br />

did permit easier jet penetration into this ice, as expected.<br />

7.0 CONCLUSIONS<br />

A large number of cuts have now been made in fresh water ice with small to<br />

moderate scale water jets; a few have also been made in a simulated sea ice with<br />

small scale water jets. In the majority of cases a narrow, clean kerf was cut in<br />

both types of ice. However, below about 40 MPa nozzle pressure the cut consisted<br />

mostly of a wide spalled trench. The kerf was apparently produced by erosion in a<br />

ductile material while the spalled trench was apparently produced by brittle frac-<br />

ture.<br />

As the fresh water ice temperature dropped substantially below freezing,<br />

a considerable reduction in penetration capability occurred. This was apparently<br />

due to an increase in ice strength. A first approximation of this effect was<br />

obtained by applying an empirical correction factor to the penetration - jet para­<br />

meters relationship based on the estimated temperature of the ice near the surface.<br />

This factor enabled the data from the entire range of temperatures to be explained<br />

by a single highly significant regression equation.<br />

With all the fresh water ice cutting data taken together, equation (5)<br />

represents a statistically excellent fit to the data. It confirms that the jet<br />

parameters, hydraulic power and the square root of traverse speed, are the important<br />

factors and that ice strength can be taken into account by a simple empirical ice<br />

T /20<br />

temperature factor (e i ). Use of this entire body of data has also revealed<br />

that nozzle stand-off does have a statistically significant effect, albeit a small<br />

one.<br />

1132


7. Coveney, D. B.<br />

Brierley, W. H.<br />

8. Pounder, E. R.<br />

1134<br />

"Cutting Cold River Ice with Water Jets during the<br />

Winter 1978-79". LTR-LT-108, National Research Coun­<br />

cil of Canada, January 1980.<br />

"The Physics of Ice". Pergamon Press Ltd., London,<br />

England, 1965.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!