+ All Categories
Home > Documents > THE UNIVERSITY OF CALGARY Avalanche Prediction for ... · avalanches to Sam Colbeck, Bert Davis,...

THE UNIVERSITY OF CALGARY Avalanche Prediction for ... · avalanches to Sam Colbeck, Bert Davis,...

Date post: 11-Aug-2020
Category:
Upload: others
View: 3 times
Download: 0 times
Share this document with a friend
275
THE UNIVERSITY OF CALGARY Avalanche Prediction for Persistent Snow Slabs by James Bruce Jamieson A DISSERTATION SUBMITTED TO THE FACULTY OF GRADUATE STUDIES IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY DEPARTMENT OF CIVIL ENGINEERING CALGARY, ALBERTA November, 1995 James Bruce Jamieson 1995
Transcript

THE UNIVERSITY OF CALGARY

Avalanche Prediction for Persistent Snow Slabs

by

James Bruce Jamieson

A DISSERTATION

SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE

DEGREE OF DOCTOR OF PHILOSOPHY

DEPARTMENT OF CIVIL ENGINEERING

CALGARY, ALBERTA

November, 1995

James Bruce Jamieson 1995

ABSTRACT

Two field tests of snow slab stability, the shear frame test and the rutschblock test,

were studied at avalanche forecasting areas in British Columbia and Alberta during the

winters of 1992-93 to 1994-95. Field work focused on persistent weak snowpack layers

consisting of surface hoar or faceted crystals that are the failure planes for most fatal slab

avalanche accidents in Canada.

The shear frame test was refined through field and finite element studies. Effects of

different frame designs were identified. Shear strength measurements were shown to

decrease as the distance between the frame and the weak layer decreased. Field studies of

the effect of loading rate and shear frame area on shear strength confirmed previous

studies. Using different shear frame operators did not affect the resulting strength

measurements provided the operators maintained consistent technique. One particular

shape of fracture surface was associated with significantly higher strength measurements.

The strength measurements from the first two tests proved to be more variable than

measurements from subsequent tests on the same weak layer.

Shear frame stability indices for natural avalanches and for skier-triggered dry slab

avalanches were refined by incorporating an adjustment for normal load that depended on

microstructure of the weak layer. The stability index for skier triggering was further

refined by adjusting for the distance the skis penetrate the snow surface. Skier stability

indices based on shear frames tests at both avalanche slopes and safe study sites were

correlated with skier-triggered dry slab avalanches. When compared with other forecasting

variables, the skier-stability index based on study site tests ranked first or second in

predictive value.

Closely spaced rutschblocks on nine avalanche slopes were used to identify

snowpack and terrain factors that affect rutschblock results. The frequency of

skier-triggered avalanches for common rutschblock scores in the avalanche start zones

was determined and shown to be similar to a Swiss study in a different snowpack. For a

given rutschblock score, persistent slabs were triggered more frequently than

non-persistent slabs.

v

Limitations of shear frame stability indices and rutschblock tests related to slope

inclination and terrain were identified.

v

ACKNOWLEDGEMENTS

I am indebted to Colin Johnston for the advice and discussions that guided this

investigation and for reviewing the chapters of this dissertation thoroughly and quickly.

For financial support for the entire research project, I am grateful to Canada’s

Natural Sciences and Engineering Research Council, Mike Wiegele Helicopter Skiing

(MWHS), Canadian Mountain Holidays (CMH), and members of the BC Helicopter and

Snowcat Skiing Operators' Association.

For their commitment to the research project and willingness to sort out the

inevitable difficulties, my thanks to Mike Wiegele and Bob Sayer from Mike Wiegele

Helicopter Skiing, to Mark Kingsbury, Walter Bruns, Colani Bezzola, Rob Rohn and

Bruce Howatt from Canadian Mountain Holidays, to Clair Israelson, Tim Auger, Marc

Ledwidge, Gerry Israelson, Dave Skjönsberg, Bruce McMahon and Terry Willis from the

Canadian Parks Service, and to Jack Bennetto, John Tweedy, Peter Weir and Gordon

Bonwick from the BC Ministry of Transportation and Highways.

For their expertise and field work at various times during the recent winters, I am

grateful to Leanne Allison, Peter Ambler, Roger Atkins, Ken Black, James Blench, Jeff

Bodnarchuk, Alex Brunet, Andrew Bullock, Steve Chambers, Peter Clarkson, Sam

Colbeck, Aaron Cooperman, Alan Evenchick, Jamie Fennell, Sylvia Forest, Michelle

Gagnon, Will Geary, Jeff Goodrich, Sue Gould, Brian Gould, Jim Gudjonson, Todd Guyn,

Reg Hawryluk, Mike Henderson, Larry Hergot, Jim Haberl, Rob Hemming, Karsten

Heuers, Jill Hughes, Gerry Israelson, Dena Jansen, John Kelly, Troy Kirwan, Karl Klassen,

Marc Ledwidge, Garth Lemke, Janet Lohmann, Kevin Marr, Greg McAuley, Rod

McGowan, Tony Moore, Al McDonald, Bruce McMahon, Derek Peterson, Cathy Ross,

Ken Schroeder, Lisa Palmer, Simon Parboosingh, Lisa Richardson, Peter Schaerer, John

Schleiss, Mark Shubin, Bert Skrypnyk, Dave Smith, Alex Taylor, Ty Trand, Julie

Timmins, John Tweedy, Scott Ward, Rupert Wedgewood, George Weetman, Barry

Widas, Terry Willis, Adrian Wilson, Percy Woods, Chris Worobets, Kobi Wyss, and Linda

Zurkirchen. My apologies to anyone I may have omitted.

vii

My thanks for helpful discussions on field work, the mountain snowpack and

avalanches to Sam Colbeck, Bert Davis, Paul Föhn, Jill Hughes, Clair Israelson, Gerry

Israelson, Dave McClung, Ron Perla, Peter Schaerer, Chris Stethem, Martin Schneebeli,

Jürg Schweizer and the guides at Canadian Mountain Holidays and Mike Wiegele

Helicopter Skiing.

Jill Hughes helped compile the data. Peter Schaerer, Jürg Schweizer and Alaa Sherif

each paraphrased sections of papers from German. Bert Davis got me interested in

classification trees and provided useful advice on Chapter 9. Martin Schneebeli provided

helpful comments on Chapters 5 and 7. Julie Lockhart proofread the entire manuscript.

Thanks to Chris Stethem for the photo of Ron Perla at the cracked bed surface in

Chapter 8, and to Jill Hughes and Mark Shubin for the photos of snowpack tests in

Chapter 1.

During this project, I was encouraged by many people including Alan Dennis, Jim

Bay, Jack Bennetto, Colani Bezzola, Bob Day, Phil Hein, Clair Israelson, Brian Langan,

John Morrall, Chris Stethem, Adrian Wilson, Jackie Wilson, my family and especially Julie

Lockhart.

My thanks to all who contributed to, or supported, this endeavour.

vii

TABLE OF CONTENTSApproval Page . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii. . .

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v. . . .

Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii. . .

Table of Contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix. . .

List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiii. .

List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xv. . .

List of Symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xxi. .

1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1. . .1.1 Effects of Avalanches . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1. . . .

1.2 Avalanche Hazard Mitigation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2. . . .

1.3 Mountain Snowpack . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3. . . .

1.4 Snow Metamorphism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4. . . .

1.5 Failure of Snow Slopes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8. . . .

1.6 Weak Snowpack Layers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9. . . .

1.7 Avalanche Forecasting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11. . .

1.8 Computer Assisted Forecasting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16. . .

1.9 Atypical Snowpack Characteristics of Accident Avalanches . . . . . . . . . . . . . . 17. . .

1.10 Skier-Triggering of Persistent Weak Layers . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18. . .

1.11 Snow Profiles and Snowpack Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19. . .

1.12 Objective and Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22. . .

2 LITERATURE REVIEW . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25.2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25. . .

2.2 Slab Failure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25. . .

2.3 Shear Frame . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27. . .

2.4 Slope-Specific Stability Indices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32. . .

2.5 Extrapolated Stability Indices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36. . .

2.6 Flaws in Weak Layers and Spatial Variability of Stability Indices . . . . . . . . . 39. . .

2.7 Rutschblock . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41. . .

2.8 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44. . .

3 METHODS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47.3.1 Study Areas and Co-operating Organizations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47. . .

3.2 Sites for Snowpack Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49. . .

ix

Table of Contents, continued3.3 Equipment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51. . .

3.4 Measurement of Slab Weight per Unit Area . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53. . .

3.5 Shear Frame Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54. . .

3.6 Rutschblock Test . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57. . .

3.7 Comparison of Rutschblock and Shear Frame Tests . . . . . . . . . . . . . . . . . . . . . . 59. . .

3.8 Avalanche Activity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61. . .

4 FIELD STUDIES OF THE SHEAR FRAME TEST . . . . . . . . . . . . . . . . 65.4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65. . .

4.2 Statistical Distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65. . .

4.3 Variability and Number of Tests for Required Precision . . . . . . . . . . . . . . . . . . 69. . .

4.4 Fracture Surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71. . .

4.5 Loading Rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73. . .

4.6 Test Sequence Variability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75. . .

4.7 Effect of Delay . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76. . .

4.8 Frame Placement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77. . .

4.9 Variability Between Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79. . .

4.10 Size Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81. . .

4.11 Effect of Normal Load . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83. . .

4.12 Frame Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86. . .

4.13 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92. . .

5 FINITE ELEMENT STUDIES OF THE SHEAR FRAME TEST . . . 95.5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95. . .

5.2 The Model and Assumptions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95. . .

5.3 Basic Stress Distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98. . .

5.4 Effect of Frame Placement on Stress Distribution . . . . . . . . . . . . . . . . . . . . . . . . 100.

5.5 Effect of Frames Placed in Hard and Soft Slabs . . . . . . . . . . . . . . . . . . . . . . . . . . 101.

5.6 Effect of Spacing Between Cross-members . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102.

5.7 Effect of Cross-Member Height . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103.

5.8 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105.

6 SHEAR FRAME RESULTS AND STABILITY INDICES . . . . . . . . . 1076.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107.

6.2 Shear Strength of Weak Layers Related to Density . . . . . . . . . . . . . . . . . . . . . . . 107.

ix

Table of Contents, continued6.3 Shear Strength of Weak Layers Related to Hand Hardness . . . . . . . . . . . . . . . 114.

6.4 Characteristics of Persistent Slab Avalanches . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116.

6.5 Predicting Natural Avalanches on Test Slopes . . . . . . . . . . . . . . . . . . . . . . . . . . . 121.

6.6 Predicting Natural Avalanches of Persistent Slabs on Surrounding Slopes 125.

6.7 Predicting Skier-Triggered Avalanches on Test Slopes . . . . . . . . . . . . . . . . . . . 138.

6.8 A Skier Stability Index for Soft Slabs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141.

6.9 Predicting Skier-Triggered Avalanches on Surrounding Slopes . . . . . . . . . . . 147.

6.10 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156.

7 RUTSCHBLOCK RESULTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1597.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159.

7.2 Site Selection and Rutschblock Variability on Test Slopes . . . . . . . . . . . . . . . . 159.

7.3 Rutschblocks on Skier-Tested Avalanche Slopes . . . . . . . . . . . . . . . . . . . . . . . . . 169.

7.4 Relationship Between Rutschblock Scores and SK from Adjacent Shear Frame Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173.

7.5 Estimating Daniels Strength from Rutschblock Scores . . . . . . . . . . . . . . . . . . 178.

7.6 Relating Rutschblock Scores to Skier-Triggered Dry Slabs inSurrounding Terrain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180.

7.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182.

8 FALSE STABLE PREDICTIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1858.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185.

8.2 Case Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185.

8.3 Characteristics Associated with False Stable Predictions . . . . . . . . . . . . . . . . . 190.

8.4 Remote Triggering and Transitional Stability for SK . . . . . . . . . . . . . . . . . . . . . . 190.

8.5 An Alternative Failure Mode for Primary Fractures . . . . . . . . . . . . . . . . . . . . . . 191.

8.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193.

9 APPLICATIONS OF SHEAR FRAME STABILITY INDICES TOAVALANCHE FORECASTING . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1959.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195.

9.2 Forecasting Variables for Natural Avalanches of Persistent Slabs . . . . . . . . . 198.

9.3 A Multivariate Forecasting Model for Natural Avalanches InvolvingPersistent Slabs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202.

9.4 Forecasting Variables for Skier-Triggered Avalanches of Persistent Slabs . 211.

xi

Table of Contents, concluded9.5 A Multivariate Forecasting Model for Skier-Triggered Avalanches

Involving Persistent Slabs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214.

9.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220.

10 CONCLUSIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22310.1 Field and Finite Element Studies of the Shear Frame Test . . . . . . . . . . . . . . 223.

10.2 Shear Strength of Weak Layers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225.

10.3 Shear Frame Stability Indices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225.

10.4 Rutschblock Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227.

10.5 False Stable Predictions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228.

11 RECOMMENDATIONS FOR FURTHER RESEARCH . . . . . . . . . . 229

REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231

A ESTIMATING DENSITY FROM MICROSTRUCTURE . . . . . . . . 247A.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247.

A.2 Hand Hardness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247.

A.3 Mean Densities by Microstructure and Hand Hardness . . . . . . . . . . . . . . . . . . 247.

B ERROR ANALYSIS FOR STABILITY INDICES . . . . . . . . . . . . . . . . 253.

B.1 Sources of Variability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253.

B.2 Variability for Index SN . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254.

B.3 Variability for Index SK . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255.

C EXAMPLE OF FIELD NOTES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257

xi

LIST OF TABLES

No. Title Page1.1 Forecasting Example Using Simplified Checklist . . . . . . . . . . . . . . . . . . . . . . . . . . 14. . .

2.1 Possibilities for Primary Fracture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26. . .

2.2 Field Studies of Shear Frame Stability Indices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33. . .

3.1 Study Sites and Locations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48. . .

3.2 Avalanche Size Classification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63. . .

4.1 Normality of Large Sets of Shear Frame Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66. . .

4.2 Number of Shear Frame Tests for Required Precision . . . . . . . . . . . . . . . . . . . . . 71. . .

4.3 Assessment of Common Shapes of Fracture Surfaces . . . . . . . . . . . . . . . . . . . . . . 72. . .

4.4 Mean Shear Strength for Various Loading Times . . . . . . . . . . . . . . . . . . . . . . . . . . 74. . .

4.5 Effect of Test Sequence on Variability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76. . .

4.6 Effect of Delay on Shear Strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77. . .

4.7 Effect of Frame Placement on Shear Strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78. . .

4.8 Effect of Different Operators on Shear Strength . . . . . . . . . . . . . . . . . . . . . . . . . . . 80. . .

4.9 Effect of Shear Frame Area on Mean Strength and Variance . . . . . . . . . . . . . . . 82. . .

4.10 Effect of Normal Load on the Daniels Strength . . . . . . . . . . . . . . . . . . . . . . . . . . . 84. . .

4.11 Shear Frame Specifications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87. . .

4.12 Effect of Shear Frame Design on Mean Strength . . . . . . . . . . . . . . . . . . . . . . . . . . 89. . .

5.1 Material Properties for Finite Element Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96. . .

5.2 Finite Element Models of the Shear Frame Test . . . . . . . . . . . . . . . . . . . . . . . . . . . 96. . .

6.1 Strength-Density Regressions by Microstructure . . . . . . . . . . . . . . . . . . . . . . . . . . 109. .

6.2 Comparison of Avalanche Reports and Investigations from Columbia Mountains 1990-95 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117. .

6.3 Characteristics of Investigated Dry Slab Avalanches from ColumbiaMountains 1990-95 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118. .

6.4 Percentage of Slabs that Failed for Skier Stability Indices . . . . . . . . . . . . . . . . . 145. .

7.1 Skier Stability Index SK from Shear Frame Tests Adjacent to RutschblockTests for Persistent and Non-Persistent Microstructures . . . . . . . . . . . . . . . . . . . 175. .

7.2 Skier Stability Index SK from Shear Frame Tests Adjacent to RutschblockTests on Slopes of at Least 20° . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177. .

7.3 Relative Frequency of Skier-Triggered Dry Slab Avalanches inSurrounding Terrain within one Day of Rutschblock Tests on Study Slope . 181. .

xiii

List of Tables, concluded8.1 False Stable Predictions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191. .

9.1 Spearman Rank Correlations Between Forecasting Variables and the DailyMaximum Size of Natural Avalanches Involving Persistent Slabs . . . . . . . . . . 199. .

9.2 Classification Trees for Daily Maximum Size of Natural Avalanches ofPersistent Slabs in the Purcell Mountains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206. .

9.3 Contingency Table for Natural Slab Avalanches Involving Persistent Slabsin the Purcell Mountains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207. .

9.4 Classification Trees for Daily Maximum Size of Natural Avalanches ofPersistent Slabs in the Cariboos and Monashees . . . . . . . . . . . . . . . . . . . . . . . . . . . 209. .

9.5 Contingency Table for Daily Maximum Size of Natural Avalanches ofPersistent Slabs in the Cariboo and Monashee Mountains . . . . . . . . . . . . . . . . . . 210. .

9.6 Spearman Rank Correlations Between Forecasting Variables and the DailyMaximum Size of a Skier-Triggered Persistent Slab . . . . . . . . . . . . . . . . . . . . . . . 212. .

9.7 Classification Trees for Daily Maximum Size of Skier-Triggered PersistentSlabs in the Cariboos and Monashees, 1992-93 to 1994-95. . . . . . . . . . . . . . . . 216. .

9.8 Contingency Table for Daily Maximum Size of Skier-Triggered PersistentSlabs in Cariboo and Monashee Mountains, 1992-93 to 1994-95. . . . . . . . . . . 217. .

9.9 Classification Trees Results for Daily Maximum Size of Skier-TriggeredPersistent Slabs in the Purcell Mountains, 1992-93 to 1994-95. . . . . . . . . . . . . 218. .

9.10 Contingency Table for Daily Maximum Size of Skier-Triggered PersistentSlab in Purcell Mountains, 1992-93 to 1994-95. . . . . . . . . . . . . . . . . . . . . . . . . . . . 220. .

A.1 Density of Layers Grouped by Hand Hardness and Microstructure . . . . . . . . 248. .

A.2 Regression Parameters for Estimating Density from Resistance andMicrostructure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251. .

xiii

LIST OF FIGURES

No. Title Page1.1 Avalanche fatalities in Canada, 1980-1995. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1. . . . .

1.2 Surface hoar on tree and snow surface. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4. . . . .

1.3 Rounding metamorphism. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5. . . . .

1.4 Faceting metamorphism. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6. . . . .

1.5 Avalanche path consisting of start zone, track and runout. . . . . . . . . . . . . . . . 8. . . . .

1.6 A point release avalanche. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8. . . . .

1.7 A small slab avalanche showing the crown fracture. . . . . . . . . . . . . . . . . . . . . . . 9. . . . .

1.8 Recently deposited snow layers including a thick weak layer of lowdensity snow and a thin weak layer of surface hoar. . . . . . . . . . . . . . . . . . . . . . . 9. . . . .

1.9 Crown fracture showing failure plane of surface hoar at base of slab. . . . . . 10. . . .

1.10 Importance of forecasting data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13. . . .

1.11 Microstructure of failure plane for fatal slab avalanche accidents inCanada, 1972-91 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18. . . .

1.12 Testing hand hardness during a snow profile observation. . . . . . . . . . . . . . . . . 19. . . .

1.13 The shovel shear test used primarily to identify weak snowpack layers. . . . 20. . . .

1.14 Compression test. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20. . . .

1.15 Shear frame test . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21. . . .

1.16 Rutschblock test showing displaced block. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22. . . .

2.1 Slab nomenclature. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26. . . .

2.2 Shear frame showing rear cross-member and two intermediatecross-members that distribute the load . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29. . . .

2.3 Underside view of finger-fin frame. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30. . . .

2.4 Cross-section of slab showing location of peak shear stress induced bystatic skier. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35. . . .

2.5 Avalanche activity and concurrent values of S35 from Cariboo andMonashee Mountains, 1990-92. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38. . . .

2.6 Rutschblock test . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41. . . .

2.7 Percentage of slab avalanches and concurrent rutschblock scores . . . . . . . . . 42. . . .

2.8 Rutschkeil test . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42. . . .

2.9 Cord-cut rutschblock . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43. . . .

3.1 Location of study sites and mountain ranges. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47. . . .

xv

List of Figures, continued3.2 Mt. St. Anne Study Plot at 1900 m in the Cariboo Mountains. . . . . . . . . . . . 49. . . .

3.3 Field staff approaching a small slab avalanche. . . . . . . . . . . . . . . . . . . . . . . . . . . . 50. . . .

3.4 Field staff approach a 1.6 m crown fracture for profiles and stability tests. 51. . . .

3.5 Equipment used for shear frame tests and measurement of slab weightper unit area. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52. . . .

3.6 Rutschblock saws. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54. . . .

3.7 Shear frame test. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56. . . .

3.8 Rutschblock isolated on the sides by shovelled trenches. . . . . . . . . . . . . . . . . . 57. . . .

3.9 Rutschblock isolated on the sides and upper wall by cord cutting. . . . . . . . . 57. . . .

3.10 Rutschblock test showing displaced block. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59. . . .

4.1 Shapiro-Wilk test for normality for 28 sets of shear frame data. . . . . . . . . . . 67. . . .

4.2 Frequency distributions for 8 sets of shear frame tests for which p < 0.05from Shapiro-Wilk test for normality. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68. . . .

4.3 Frequency distribution for coefficients of variation of shear strength. . . . . . 70. . . .

4.4 Effect of loading time on shear strength for 10 experiments with variousmanual loading rates. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74. . . .

4.5 Effect of sequence number on standard deviation. . . . . . . . . . . . . . . . . . . . . . . . 76. . . .

4.6 Effect of normal load on strength from previous studies. . . . . . . . . . . . . . . . . 84. . . .

4.7 Measured and predicted effect of normal load on Daniels strength. . . . . . . . 85. . . .

4.8 Shear frames used for comparative studies of frame design and sizeeffects. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86. . . .

4.9 Twelve strength comparisons of short frame with standard frame. . . . . . . . 90. . . .

5.1 Geometry and loading for finite element model of standard shear frameplaced 3 mm above weak layer. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95. . . .

5.2 Finite element mesh for snow in left compartment and underlying weaklayer and substratum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98. . . .

5.3 Stress contours for σxz for standard frame placed in soft superstratum3 mm above weak layer. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99. . . .

5.4 Shear stress σXZ in weak layer for standard frame placed 3 mm aboveweak layer representing average σXZ of 1.0 kPa. . . . . . . . . . . . . . . . . . . . . . . . . . 99. . . .

5.5 Shear stress σXZ in weak layer for standard frame placed in weak layerand 3 mm above weak layer. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100. . .

5.6 Distribution of σXZ for the standard frame placed in soft and hardsuperstrata. In both cases, the frame is 3 mm above the weak layer. . . . . . . 101. . .

List of Figures, continued

xv

5.7 Distribution of σXZ for 5-cross-member and standard frame. . . . . . . . . . . . . . . 102. . .

5.8 Distribution of σXZ in weak layer for standard and short frames placed3 mm above the weak layer. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104. . .

5.9 Distribution of σXZ in weak layer for standard and short frames placed 1mm into the weak layer. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104. . .

6.1 Daniels strength for weak layers by microstructure and density. . . . . . . . . . . 110. . .

6.2 Normalized regression variance for Group I and II microstructures. . . . . . . 111. . .

6.3 Shear strengths from present study compared with those from Perla andothers (1982) for four common microstructures. . . . . . . . . . . . . . . . . . . . . . . . . . 113. . .

6.4 Shear strength by hand hardness for common microstructures. . . . . . . . . . . . 114. . .

6.5 Shear strength plotted against scaled hand hardness for decomposed andfragmented particles and for faceted grains. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116. . .

6.6 Cross section of typical dry slab avalanches. Layer thicknesses aremeasured vertically. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119. . .

6.7 Relative frequency of microstructures for superstratum, weak layer andsubstratum of dry slab avalanches in Columbia Mountains, 1990-95. . . . . . 119. . .

6.8 Resistance for superstratum, weak layer and substratum of dry slabavalanches in Columbia Mountains, 1990-95. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121. . .

6.9 Values of SN for slopes that did and did not avalanche naturally. . . . . . . . . . 122. . .

6.10 Stability trend for natural avalanches on surface hoar buried 19 January1993. in the Purcell Mountains. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130. . .

6.11 Stability trend for surface hoar layer in the Purcell Mountains buried 10February 1993. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131. . .

6.12 Stability trend for a layer of surface hoar buried 6 February 1994 in thePurcell Mountains. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132. . .

6.13 Stability trend for a layer of surface hoar buried 29 December 1993 in theCariboo and Monashee Mountains near Blue River, BC. . . . . . . . . . . . . . . . . . 133. . .

6.14 Stability trend for a layer of surface hoar buried 5 February 1994 in theCariboo and Monashee Mountains near Blue River, BC. . . . . . . . . . . . . . . . . . 134. . .

6.15 Stability trend for a layer of surface hoar buried 7 January 1995 in theCariboo and Monashee Mountains near Blue River, BC. . . . . . . . . . . . . . . . . . 135. . .

6.16 Stability trend for a layer of facets formed in October 1993 in JasperNational Park. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136. . .

6.17 Stability trend for a layer of surface hoar buried 8 February 1994 inJasper National Park. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137. . .

6.18 Stability index SS for skier-tested avalanche slopes. . . . . . . . . . . . . . . . . . . . . . . 139. . .

List of Figures, continued

xvii

6.19 Effect of ski penetration on skier-induced stress. . . . . . . . . . . . . . . . . . . . . . . . . . 142. . .

6.20 Profiles of averaged densities for high and low density slabs from theColumbia Mountains. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143. . .

6.21 Skiing penetration for mean slab density and estimated density at 0.3 m . . 144. . .

6.22 Skier stability index SK for skier-tested avalanche slopes. . . . . . . . . . . . . . . . . . 146. . .

6.23 Skier stability trend for surface hoar layer buried 18 January 1993 in thePurcell Mountains. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149. . .

6.24 Skier stability trend for surface hoar layer buried 10 February 1993 inthe Purcell Mountains. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149. . .

6.25 Skier stability trend for surface hoar layer buried on 6 February 1994 inthe Purcell Mountains. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150. . .

6.26 Skier stability trend for surface hoar layer buried 7 January 1995 in thePurcell Mountains. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151. . .

6.27 Skier stability trend for surface hoar layer buried 6 February 1995 in thePurcell Mountains. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152. . .

6.28 Skier stability trend for surface hoar layer buried 10 February 1993 in theCariboos and Monashees near Blue River, BC. . . . . . . . . . . . . . . . . . . . . . . . . . . 153. . .

6.29 Skier stability trend for the surface hoar layer buried 29 December 1993in the Cariboo and Monashee Mountains near Blue River, BC. . . . . . . . . . . . 154. . .

6.30 Skier stability trend for the surface hoar layer buried 5 February 1994 inthe Cariboo and Monashee Mountains near Blue River, BC. . . . . . . . . . . . . . 155. . .

6.31 Skier stability trend for the surface hoar layer buried 7 January 1995 inthe Cariboo and Monashee Mountains near Blue River, BC. . . . . . . . . . . . . . 156. . .

7.1 Rutschblock scores from Mt. St. Anne in the Cariboo Mountains, eastaspect, 1900 m on 13 February 1991. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160. . .

7.2 Rutschblock scores from a northwest facing slope in Miledge valley inCariboo Mountains on 6 March 1991. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161. . .

7.3 Rutschblock scores from Mt. St. Anne in the Cariboo Mountains, northaspect, 1900 m on 6 April 1991. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162. . .

7.4 Rutschblock scores from a northeast-facing slope in Miledge valley in theCariboo Mountains on 7 January 1992. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163. . .

7.5 Rutschblock scores from a north-facing slope in Miledge valley in theCariboo Mountains on 19 January 1992. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164. . .

7.6 Rutschblock scores from a north-facing slope in Miledge valley in theCariboo Mountains on 3 February 1992. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165. . .

xvii

List of Figures, continued7.7 Rutschblock scores on Mt. St. Anne in the Cariboo Mountains, northeast

aspect, 1900 m on 29 February 1992. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166. . .

7.8 Rutschblock scores on a northeast-facing slope in the MonasheeMountains on 31 March 1992. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167. . .

7.9 Rutschblock scores on a northeast-facing slope in the MonasheeMountains on 7 April 1992. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168. . .

7.10 Relative frequency of skier-triggered slabs on skier-tested avalancheslopes from Föhn (1987b) and present study. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170. . .

7.11 Relative frequency of skier-triggering for persistent and non-persistentslabs on skier-tested avalanche slopes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171. . .

7.12 Slab thicknesses of persistent and non-persistent skier-tested slabs. . . . . . . 173. . .

7.13 Normalized deviations of SK from mean values for particular rutschblockscores. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176. . .

7.14 Mean, standard deviation and standard error for median rutschblockscores from adjacent tests. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178. . .

7.15 Daniels strengths estimated from rutschblock scores plotted againstmeasured Daniels strengths from adjacent shear frame tests. . . . . . . . . . . . . . 179. . .

7.16 Relative frequency of one or more skier-triggered avalanches insurrounding terrain within one day of study-slope rutschblock results . . . . 182. . .

8.1 Cross-section of test site, crown fracture and bed surface at slabavalanche on Mt. Albreda in the Monashee Mountains . . . . . . . . . . . . . . . . . . 186. . .

8.2 Remotely triggered slab avalanche in the Purcell Mountains on 24February 1994. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188. . .

8.3 Cross-sections of snowpack at trigger point, profile site on propagationpath and crown for a remotely triggered slab avalanche . . . . . . . . . . . . . . . . . . 188. . .

8.4 Cracks in bed surface at Whistler Mountain, February 1979. . . . . . . . . . . . . 193. . .

9.1 Box plots of the daily maximum size of natural avalanche involving apersistent slab against various forecasting variables . . . . . . . . . . . . . . . . . . . . . . 201. . .

9.2 Classification tree for daily maximum size of natural avalanches ofpersistent slabs in the Purcell Mountains using forecasting variables butexcluding SN38. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205. . .

9.3 Classification tree for the daily maximum size of natural avalanches ofpersistent slabs in the Purcell Mountains based on data from the wintersof 1992-93 to 1994-95. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206. . .

9.4 Classification tree for the daily maximum size of natural avalanches ofpersistent slabs in the Cariboo and Monashee Mountains based on 150days from the winters of 1992-93 to 1994-95. . . . . . . . . . . . . . . . . . . . . . . . . . . 210. . .

xix

List of Figures, concluded9.5 Box plots of the daily maximum size of a skier-triggered persistent slab

against various forecasting variables showing median (small rectangle),lower and upper quartiles (box) and minima and maxima (whiskers). . . . . 213. . .

9.6 Classification tree for the daily maximum size of skier-triggeredpersistent slab in the Cariboo and Monashee Mountains based on datafrom the winters of 1992-93 to 1994-95. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215. . .

9.7 Classification tree for the daily maximum size of skier-triggered persistent slabs in the Purcell Mountains using meteorological forecastingvariables but excluding SK38. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218. . .

9.8 Classification tree for the daily maximum size of skier-triggered persistent slabs in the Purcell Mountains using meteorological forecastingvariables and including SK38. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219. . .

A.1 Density by hand hardness for six common classes of microstructure. . . . . . 249. . .

C.1 Example of field notes for profile, shear frame tests and rutschblock test. . 258. . .

xix

LIST OF SYMBOLS

αmaxangle between snow surface and peak shear stress due to skier, etc.

λ fractional settlement

φ normal load adjustment for shear strength

ρ average slab density

ρ0 estimated snow density at surface

ρ30 estimated snow density at 0.30 m below surface

ρicedensity of ice (917 kg/m3)

σ stress

σV vertical stress due to overburden

σXZ shear stress parallel to snow surface

∆σXZ shear stress parallel to snow surface due to artificial load such as a skier

∆σ'XZ shear stress parallel to snow surface due to a skier, adjusted for ski penetration

σZZ normal stress perpendicular to snow surface

Σ shear strength

Σ100 shear strength measured with a 0.01 m2 shear frame

Σ250 shear strength measured with a 0.025 m2 shear frame

Σφshear strength adjusted for normal load

Σ∞Daniels strength (shear strength of an arbitrarily large specimen)

Σ∞* Daniels strength estimated from rutschblock score

Ψ angle between snow surface and horizontal

A, a, B empirical constants

b shear frame width

d shear frame depth

df degrees of freedom

D difference in strength measurements

F F statistic

g acceleration due gravity (9.81 m/s2)

h slab thickness (measured vertically)

List of Symbols, continued

xxi

hSZ slab thickness in start zone

HN height of snowfall between consecutive morning weather observations

HNW water equivalent of height of precipitation (ran and snowfall) betweenconsecutive morning weather observations

HS height of snowpack (measured vertically)

HST height of storm snowfall (measured vertically)

HSTW water equivalent of height of presipitation during storm

L line load due to skier

MN1 size class of largest natural slab avalanche from previous day

MN2 sum of size classes of largest natural slab avalanche from previous two days

MS1 size class of largest skier-triggered slab avalanche from previous day

MS2 sum of size classes of largest skier-triggered slab avalanche from previous twodays

MxN size class of largest natural slab avalanche on forecast day

MxS size class of largest skier-triggered slab avalanche on forecast day

n, N number of data

p probability associated with a statistic, significance level

P precision expressed as a fraction of the mean

PB barometric pressure

PF foot penetration

PK average of ski penetration while standing and after two jumps on same spot,estiamte of maximum penetration of skis during skiing

r correlation coefficient

R Spearman rank correlation coefficient

R2 coefficient of determination

RH relative humidity

s standard deviation

se standard error (standard deviation of the mean)

S stability index for natural slab avalanches calculated for a specific start zone,includes frame size adjustment and normal load adjustment for granular snow

S35 stability index for natural slab avalanches calculated for 35° slopes, includesframe size adjustment and normal load adjustment for granular snow

S' stability index for slab avalanches triggered by a skier, etc., includes frame sizeadjustment and normal load adjustment for granular snow

List of Symbols, concluded

xxi

SN stability index for natural avalanches calculated for a specific start zone,includes frame size adjustment and normal load adjustment for granular snow

SN38 stability index for natural slab avalanches calculated for 38° slopes, includesframe size adjustment and microstructure-dependent effect of normal load onweak layer

SK stability index for skier-triggered slab avalanches calculated for specific startzone, includes adjustments for ski penetration, frame size andmicrostructure-dependent effect of normal load on weak layer

SK* mean value of SK for a particular rutschblock score

SK38 stability index for skier-triggered slab avalanches calculated for 38° slopes,includes adjustments for ski penetration, frame size and microstructure-dependent effect of normal load on weak layer

t students t-statistic

T classification tree

T' classification sub-tree

Ta air temperature at time of weather observation

Tmin minimum air temperature in 24 hours prior to morning weather observation

Tmax maximum air temperature in 24 hours prior to morning weather observation

u standard normal variable

v vertical co-ordinate measured downwards from snow surface

V coefficient of variation

w distance between shear frame cross-members

x downslope co-ordinate (parallel to snow surface)

y cross-slope co-ordinate

z slope-perpendicular co-ordinate measured downwards from snow surface

xxiii

1 INTRODUCTION

1.1 Effects of Avalanches

In Canada, most avalanches have no effect on people, structures or roads. The vast

majority start in the backcountry without human involvement and come to rest without

encountering people or human artefacts. Only when avalanches have the potential to affect

people, structures or transportation facilities is there a hazard.

Avalanches and closures for avalanche control delay traffic on highways and the cost

of such delays is substantial. Morrall and Abdelwahab (1992) estimate the cost of a two

hour closure at Rogers Pass at $50,000 to $90,000 depending on the proportion of heavy

vehicles in a traffic volume of 350 vehicles/h in each direction. Blattenberger and Fowles

(1995) estimate the cost of a one day closure of the Little Cottonwood Highway in Utah

at US$1,410,370. The annual cost of traffic delays due to avalanche hazards far exceeds

the cost of property damaged by avalanches which averaged less than $350,000 per year in

Canada during the period 1970 to 1985 (Schaerer, 1987, p. 6).

During the years 1972 to 1995, avalanche fatalities in Canada averaged eight per

year and increased gradually as indicated by the five-year moving average in Figure 1.1.

This dissertation

focuses on predicting

avalanche hazards to

backcountry

recreationists who

account for

approximately 96% of

the 127 fatalities in

Figure 1.1, the

remainder occurring

in residential areas or

transportation

corridors.

1

Figure 1.1 Avalanche Fatalities in Canada, 1980-1995.(Schaerer, 1987; Canadian Avalanche Centre)

1.2 Avalanche Hazard Mitigation

Avalanche hazards to structures, dwellings and transportation corridors are often

mitigated by zoning—placing the elements at risk in zones where avalanche return

intervals are acceptably long or expected impact pressures are sufficiently small. When this

is not adequate, the hazard can be further mitigated by

supporting structures in avalanche start zones designed to prevent most avalanches

from starting, costing approximately $1,000,000 per hectare (Mears, 1992,

p. 44-45) and used mainly in Europe where population centres occur in mountain

areas and in Japan where use of explosives is very restricted,

defence structures usually designed to divert avalanches, such as splitting wedges,

diversion berms and snow-sheds over highways or rail lines, or to slow avalanches,

such as retarding mounds (McClung and Schaerer, 1993, p. 225-234), and

avalanche forecasting and control programs often involving closures of affected

areas during which avalanches are released by explosives.

In Canada, diversion berms are used for some minor highways but major reinforced

concrete defence structures such as snow-sheds are only economically justifiable for major

highways such as the Trans-Canada Highway through Rogers Pass.

Avalanche forecasting and control programs for transportation corridors typically

close the corridor for periods of less than a few hours during which unstable snow is

released with explosives, often delivered from helicopters or from road level with artillery.

Roads with low traffic volumes sometimes remain closed for a day or more while storm

snow stabilizes naturally, or control teams wait for suitable weather before placing

explosives from helicopters.

Avalanche hazards to lift-based ski areas are also managed with forecasting and

control programs. When control is deemed necessary by the forecaster, many areas are

stabilized with explosives before the ski area opens for the day; other areas may remain

closed for part or all of the day while slopes are stabilized. Explosives are often placed

from the ground or a helicopter, or by an “avalauncher” which uses pressurized gas, a

2

pressure vessel and barrel to propel explosives to avalanche start zones usually less than

1 km away.

Backcountry recreational operations including helicopter skiing, snowcat skiing and

ski touring manage avalanche hazards primarily by avoiding terrain where and when

hazardous avalanches are probable. Since such operations use vast mountain areas

(1,000-6,000 km2), explosives are not used, or are limited to testing for unstable snow and

to stabilizing a few selected slopes. When large avalanches are not expected, selected

slopes are often test-skied by ski guides, releasing unstable snow as small avalanches.

In Canada, avalanche forecasting is common to hazard management programs for

railways, highways, lift-based ski areas, commercial backcountry recreational operations

and parks. Avalanche forecasting is introduced in Section 1.7 following sections on the

mountain snowpack, snow metamorphism, weak snow layers and failure of snow slopes.

1.3 Mountain Snowpack

Snow crystals form in the atmosphere either by sublimation of water vapour, initially

onto small particles (freezing nuclei), and/or by accretion of super-cooled water droplets

(rime). Variations in temperature and supersatuation within the atmosphere result in

different shapes of crystals which, when precipitated, are often recognizable as different

layers of precipitation particles (Class PP). Stellar crystals with six arms are a common

form. When rime obscures the original form of precipitation particles, the resulting grains

are called graupel. When not broken by wind, precipitation particles are often 1-6 mm in

length or diameter. When first deposited, dry layers of unbroken crystals typically have a

density of 40-120 kg/m3. Alternatively, turbulence may fragment the crystals and deposit

very fine particles (< 1 mm) into relatively dense layers (> 140 kg/m3).

During generally clear weather, water vapour may sublime as surface hoar (frost)

directly onto objects and the snow surface (Figure 1.2). When buried by subsequent

snowfalls, layers of surface hoar sometimes remain weak for a month or more, often

This dissertation uses the classification of the Canadian Avalanche Association

(CAA, 1995) which closely follows the definitions of Colbeck and others (1990).

3

playing an important role in

avalanche formation.

The properties of new

snow layers vary over mountain

terrain. A particular snow storm

may deposit a dry layer at higher

elevations and a wet layer at

lower elevations where the

ambient temperature was

warmer. Commonly, more snow

falls at higher elevations than at

lower elevations. Alternatively, a

single snow storm may deposit a

denser wind-packed layer at

higher elevations than at lower

elevations where calmer

conditions prevailed. Further,

wind may deposit little or no

snow on windward slopes (and

sometimes remove previously

deposited snow from windward

slopes), depositing much of the

snow on leeward slopes.

1.4 Snow Metamorphism

Once on the ground, the properties of snow layers change over time, partly

reflecting changes in microstructure which consists of ice grains and bonds between them.

There are three main metamorphic processes that change the microstructure, and

consequently, the mechanical properties of layers: rounding and faceting which occur in

4

Figure 1.2 Surface hoar on tree and snow surface.

dry snow (≤ 0°C) and melt-freeze metamorphism that occurs when the snow temperature

cycles above and below 0°C.

1.4.1 Rounding

In the absence of a sufficiently

strong temperature gradient, particles

reduce their specific surface, primarily

by ice-to-vapour-to-ice sublimation and

evolve toward more rounded

equilibrium forms (Colbeck, 1983).

Precipitation particles with initial

dendritic forms decompose into smaller

particles (Figure 1.3), followed by the

growth of the larger particles at the

expense of the smaller particles which

disappear over time. Typically, particle

convexities suffer a net loss of mass

and concavities a net gain. Since

contact points between grains are

effectively concavities, bonds grow

between grains resulting in strengthening of the layer. Layer density also increases.

Although this is a dry snow process (≤ 0°C), the metamorphic rate increases with

temperature. In the seasonal snowpack, rounded grains (Class RG) are usually 0.2-1.0 mm

in diameter.

At an intermediate stage when the precipitation particles are still recognizable

(Figure 1.3), the particles are called partly decomposed (Colbeck and others, 1990). Such

particles are grouped with wind-broken particles into a class called decomposing and

fragmented precipitation particles (Class DF).

5

Figure 1.3 Rounding metamorphism. Watervapour sublimates as ice onto concave surfacesresulting in intergranular bonds and eventuallyin rounded equilibrium forms.

1.4.2 Faceting (Kinetic Growth)

Driven by a sufficiently strong temperature

gradient, there is a net directional flow of water

vapour, usually upwards since many layers are

warmer on top (Figure 1.4). Under these

conditions, ice generally sublimates from the top of

a crystal and water vapour deposits as ice on the

bottom of a crystal above. Larger grains grow at

the expense of smaller grains. At the intermediate

stage when flat (faceted) faces but no striations are

apparent, the grains are called faceted crystals

(Class FC). In the seasonal snowpack, faceted

crystals are typically 0.5-2 mm in size, although

smaller faceted crystals may form near the snow

surface (Colbeck and others, 1990).

When kinetic growth continues, depth hoar

(Class DH) consisting of striated crystals and subsequently skeletal forms including cup-,

column- and plate-shaped crystals are apparent (Akitaya, 1974). Crystals are often 2-6 mm

in length and may exceed 10 mm.

Prior to the formation of skeletal forms, strength usually decreases (Akitaya, 1974;

Bradley and others, 1977a, b; Adams and Brown, 1982; Armstrong, 1980, 1981) except

for relatively dense layers (> 300 kg/m3) (Akitaya, 1974; Perla and Ommanney, 1985).

The advanced skeletal forms are often stronger than the non-skeletal, striated forms.

The temperature gradient is increased and hence the faceting process is augmented

adjacent to layers of low permeability such as crusts (Seligman, 1936, p.70; Adams and

Brown, 1982; Moore, 1982; Colbeck, 1991).

1.4.3 Melt-freeze

A snow layer at 0°C may contain liquid water due to melting or rain. Under such

conditions, small grains disappear, larger grains grow rapidly until they reach 1-2 mm,

6

Figure 1.4 Faceting metamorphism.The transfer of water vapour under asufficiently strong temperaturegradient results in the progressiveshape changes.

bonding and strength decrease, and density decreases (Male, 1980). Commonly, draining

limits the liquid water content to approximately 8%. Under these conditions, the grains

arrange into clusters held together by surface tension of the liquid water. Regardless of the

liquid content, once wet snow layers freeze, they strengthen and form crusts. Since the

present study is entirely restricted to dry snow, only the crusts are of interest.

1.4.4 Effect of Temperature Gradient on the Strength of Dry Snowpacks

Rounding and faceting processes compete in dry snow layers, and layers showing

evidence of both processes are common. Faceting dominates if the temperature gradient

exceeds a critical level and rounding dominates below this level (de Quervain, 1958).

Although field workers often use 10°C/m as a rough estimate of the critical temperature

gradient (Armstrong, 1981, Colbeck, 1983), it varies considerably and is affected by the

temperature, density and permeability (Armstrong, 1980, 1981; Perla and Ommanney,

1985).

A positive temperature gradient (warmer above) is caused by diurnal warming, a

warm front or warm snow falling on a cooler snow surface. However, during much of the

winter in western Canada, positive temperature gradients are usually restricted to surface

layers and are replaced by a negative gradient (cooler above), sometimes within a few

hours and almost always within a few days, due to the upward flow of heat from the

ground toward the cooler snow surface. Also, where the snowpack thickness exceeds 1 m,

the ground-snow interface stays close to 0°C in temperate latitudes.

To consider the effect of temperature gradient and snow metamorphism on the

strength of snow layers, generalizations are necessary. Assuming an air temperature below

freezing, the magnitude of the average temperature gradient through the snowpack is

increased where the snowpack thickness is reduced, or when the air temperature is

reduced. In areas where the snowpack thickness is less than 1 m and the air temperature is

less than -10°C for an extended period, faceting often dominates and weak layers are

common. These conditions occur during early and mid-winter in many areas of the Rocky

Mountains, and in isolated areas of the Columbia Mountains, such as where the wind has

reduced the snowpack thickness. Alternatively, in most areas of the Columbia Mountains

7

where the snowpack thickness typically exceeds

2 m during mid-winter, rounding dominates,

causing most layers to strengthen. During warmer

weather in the spring, rounding and melt-freeze

processes affect most layers in both ranges. The

rounding process tends to strengthen and stabilize

the weak layers that formed during the winter.

Melt-freeze cycles alternately weaken and

strengthen layers, often diurnally.

1.5 Failure of Snow Slopes

Many, but not all, avalanche paths can be

divided into a start zone where avalanches

initiate, a track where only a few small avalanches

stop, and a runout zone where the larger

avalanches decelerate (Figure 1.5).

Avalanches release from snow slopes in two

distinct ways. When a small amount of cohesionless

snow—typically the size of a snowball—slips and

begins to slide down a slope setting additional snow

in motion, it is called a point release or loose snow

avalanche (Figure 1.6). Alternatively, when a plate

or slab of cohesive snow begins to slide as a unit

before breaking up, it is called a slab avalanche

(Figure 1.7). Following a slab avalanche, a distinct

fracture line or crown fracture is visible at the top

of the avalanche. Slab avalanches will only occur

when there is a weak layer under the cohesive layers

that make up the slab. Slab avalanches, which are

8

Figure 1.6 A point releaseavalanche.

Figure 1.5 Avalanche path consistingof start zone, track and runout.

generally larger than point

release avalanches, are the

focus of this study.

Natural avalanches start

without any human-related

trigger such as a skier, hiker,

explosive, over-snow vehicle,

etc. Avalanches triggered by

the fall of a cornice—a

sometimes massive chunk of

wind-packed snow from a

ridge top—are not considered

natural avalanches in this dissertation as in NRCC/CAA (1989) since cornice falls are

often more powerful than many human-related triggers. Only natural and skier-triggered

avalanches are considered.

1.6 Weak Snowpack Layers

In Canada, most

fatalities are caused by dry

slab avalanches (Jamieson

and Johnston, 1992a), and

the failures that release such

avalanches start and

propagate in weak layers

(e.g. McClung, 1987).

Identifying weak layers

(Figure 1.8) is fundamental

to identifying unstable slabs

so that they can be avoided.

9

Figure 1.8 Recently deposited snow layers including athick weak layer of low density snow and a thin weaklayer of surface hoar.

Figure 1.7 A small slab avalanche showing thecrown fracture.

Since weak layers are common in the Rocky Mountains throughout much of the

winter, the snowpack is generally less stable than in the Columbia Mountains (where most

of the commercial backcountry skiing in western Canada occurs).

Weak layers can either be precipitated, form on or near the snowpack surface as

surface hoar or faceted crystals, or form at depth as faceted crystals or depth hoar.

Precipitated weak layers usually consist of large stellar-, needle-, or plate-shaped crystals,

which may remain weaker and lower in density than adjacent layers during the early stages

of rounding (Figure 1.8). Such weak layers generally stabilize within a few days of

deposition and are subsequently termed non-persistent. While such layers were the failure

plane for some fatal avalanches involving amateur decision-makers with widely varying

levels of avalanche skills, Jamieson and Johnston (1992a) found no reports of fatal

avalanches involving such layers where professionals made the decisions regarding access

to avalanche terrain. In general, professionals have the techniques and experience to

manage avalanche hazards due to non-persistent weak layers.

Cooling periods as short as a single cold, clear night can produce weak layers on and

near the snow surface. Strong near-surface temperature gradients due to such cooling can

cause small faceted crystals to grow within the top 20 mm of the snowpack. In areas with

little or no wind, surface hoar crystals can grow on the snow surface. Layers of surface

hoar and/or faceted crystals that form near the surface are important in the present study

since, when buried by subsequent

snowfalls, they can remain weak

(persist) sometimes for a month or

more, and form the failure plane for

some slab avalanches (Figure 1.9).

Since these weak layers form on the

surface, they can be identified by the

date they were buried.

Prolonged periods of cold

weather, especially where the

10

Figure 1.9 Crown fracture showing failure planeof surface hoar at base of slab.

snowpack is relatively thin, can create thick layers of depth hoar and faceted crystals.

Since this faceting process is faster at the base of the snowpack where the snow

temperature is generally warmer, depth hoar is most common at, but not restricted to, the

base of the snowpack.

Weak layers of surface hoar, faceted crystals and depth hoar are termed persistent

weak layers. They accounted for the majority of avalanche fatalities in Canada from 1972

to 1992 (Jamieson and Johnston, 1992a).

The thickness of a weak layer plays an important role in slab failure. In Bader and

Salm’s (1990) slab failure model, the shear strain rates necessary for propagation are only

possible where strain is concentrated in thin weak layers. This is supported by an extensive

field study in Switzerland, where 60% of weaknesses that failed in slab avalanches or

stability tests were so thin they were classified as weak interfaces, and the remaining 40%

averaged 11 mm in thickness (Föhn, 1993). While these studies do not preclude an

important role for thick weak layers, they emphasize that slab failures often start in weak

layers so thin that they may be difficult to observe in the snowpack.

1.7 Avalanche Forecasting

Avalanche forecasts refer to the likelihood and size of avalanches, usually in terms of

avalanche hazard, avalanche danger or snow stability for areas of terrain that vary from an

entire mountain range to a specific slope. Since forecasts are constantly being refined, the

temporal extent of the forecast is usually limited. In Canada, large-scale forecasts for a

mountain range (bulletins) are usually valid for 1-7 days “unless conditions change”.

Rather than attempt to predict the likelihood of avalanches following a change in weather,

forecasters prefer to issue a new bulletin or advisory. For backcountry and lift-based

skiing operations, the forecasts, often called snow stability evaluations, are prepared each

morning.

Avalanche forecasting is inherently multivariate. Atwater (1954) proposed a list of

10 factors, some of which are quantifiable such as snowfall depth, precipitation rate and

air temperature, and some qualitative factors such as the character of the snow surface

11

prior to the storm. The exact list of factors has evolved (e.g. CAA, 1994) and varies

according to the type of forecasting operation, be it a highway, lift-based ski area or

backcountry program, and with the individual forecaster (LaChapelle, 1980). However,

the factors can be grouped by the entropy (or noise or disorder) of the information.

Avalanche activity and stability tests are low entropy information since they pertain

directly to slope stability; snowpack observations such as profiles identifying snow layers

are medium entropy information; and weather measurements that are less directly related

to slope stability are high entropy information (LaChapelle, 1980). These groups are also

called Class 1, 2 and 3, respectively (McClung and Schaerer, 1993, p. 124).

The importance of forecasting data depends not only on its entropy, but also on how

far from the relevant terrain it was obtained (Figure 1.10). In Canada, meteorological

measurements are available from

other avalanche safety operations in the same range, perhaps 100 or more km away,

by an evening exchange of weather, snowpack and avalanche information,

a weather station often within 10 to 100 km, and

sometimes from a regularly accessed study site with basic weather instruments

within 10 km of the relevant start zones.

In addition to weather measurements, snowpack observations, stability tests and

reports of avalanche activity are exchanged every evening between approximately 50

different forecasting operations in western Canada. Further, each operation records

snowpack observations, stability tests and avalanche activity within their forecast area

daily. However, backcountry operations and lift-based ski areas generally access study

sites and start zones more frequently than highway forecasting operations, particularly

those highway operations in the Coast mountains where short-term weather trends often

have a dominant effect on avalanche activity.

Terrain plays an important role in avalanche forecasting. Mesoscale avalanche

forecasts for areas typically 10-100 km across, such as parks, summarize avalanche

conditions with general reference to terrain inclination, elevation and orientation to wind

or sun. Forecasts for lift-based ski areas and highways are concerned with specific start

12

zones. Backcountry skiing operations must select terrain on the microscale with due

consideration to local terrain features since the distance between safe and unsafe terrain

may be less than 50 m. Such precise terrain selection is the most detailed application of the

same information also used by other forecasting operations. However, while local terrain

features complicate site selection and interpretation for snowpack observations and

stability tests (Chapters 7 and 8), terrain features on the same scale can be used

advantageously by ski guides to select safe routes. As shown in Figure 1.10, a simple

snowpack observation in or near an avalanche start zone such as noting that a weak layer

exists 0.4 m below the surface can be more important than an accurate measurement of

wind speed or precipitation from a weather station 10 km away.

Consider the following simplified example of the forecasting process for a

backcountry skiing operation, based loosely on the Canadian Avalanche Association’s

13

Figure 1.10 Importance of forecasting data increases with proximity to forecast area andwith decreasing entropy of information. The present study concentrates on stability testsin study sites and start zones.

14

Table 1.1 Forecasting Example Using Simplified Checklist

Factor Data Stability? Import-ance?

Confid-ence?

I Avalanche Activity and Stability Tests

AvalancheActivity

Yesterday, many slopes wereskier-tested. Two small slabs were

skier-triggered, 1 on a NE-facing slopenear treeline and 1 on a N-facing slope

above treeline.

N High High

Stability TestsA stability test resulted in a “moderate”

failure1 on a layer of surface hoar0.4 m below the surface near tree line

on an E-facing slope.

N High Low

II Snowpack Observations

ProfileA layer of 4 mm surface hoar found0.3 m below surface in study plot

profile. Overall hardness2 of slab hasincreased in last 2 days.

N High High

Settlement 0.33 m of snow from the last storm hassettled to 0.30 m in the last 2 days

? High High

Temperature buried surface hoar is at -5°C ? Low Mod.

Temp. Gradient approximately -5°C/m in slab N High Mod.

Snowpack Height 1.80 m in study plot ? Low High

II Meteorological ObservationsPrecipitation None in last 2 days Y Low High

Precipitation Rate nil Y Low High

Wind Speed averaged 20 km/h last night, nonoticeable drifting

Y High Low

Wind Direction West ? Low Low

Air Temperature Yesterday’s max. temp. was -12°C,overnight min. temp. was -19°C

? Mod. High

Weather Forecast Warming to -5°C N High Mod.

3-5 cm of snow starting this afternoon Y High Low1 “Moderate” manual force was required to induce failure. Stability tests are describedfurther in Section 1.11.2 Hardness is assessed on an ordinal scale outlined in Appendix A. The hardness of alayer or slab is relevant in various ways. It is relevant in this example because harderslabs tend to result in wider and often larger avalanches.

stability evaluation checklist (CAA, 1994) . In the morning, the forecaster makes

abbreviated notes of the relevant data for avalanche activity and stability tests (Class I

data), snowpack observations (Class II data) and meteorological observations (Class III

data) as in column two of Table 1.1. The forecaster would then mark a “Y” for the factors

that are contributing to stability, an “N” for the factors that are contributing to instability

and a “?” for factors with unclear or mixed effects. In this case, the lack of precipitation in

the last two days and the wind which was too light to cause drifting favour stability,

whereas limited avalanche activity (two small skier-triggered slabs), a stability test that

failed on a surface hoar layer, a profile in a study plot that found the surface hoar under a

stiffening slab, and forecast warming above the maximum temperature of the previous day

are indicating instability. Although the labelling of contributions to stability or instability is

simplistic, the primary value of such a checklist is in ensuring that the forecaster reviews

and assesses the various factors.

Forecasters then consider the importance of the various factors (low, moderate or

high in Column 4 of Table 1.1) for the planned skiing and the type of avalanche activity

expected. Usually informally, forecasters also assess their confidence in the factors (low,

moderate or high in Column 5 of Table 1.1). Factors in which the forecaster has high

confidence in the observations and considers important are used to select skiing terrain for

the day; in this case the avalanche activity, the profile and the forecast warming would

probably lead the forecaster to exclude large avalanche slopes above and below treeline. If

there are important factors for which the forecaster has little confidence in the available

observations, the terrain selection might be even more restricted and the forecaster would

specify field tests and observations for field work during the day to improve confidence in

some of the important factors. In this example, the forecaster can do little about the lack

of confidence in the amount of forecast snowfall, but concerns about drifting might be

reduced by observations from a ridge-top and/or helicopter during the day, and additional

profiles and stability tests near and above tree-line might be assigned to the technicians and

guides to better determine where the buried layer of surface hoar exists. Since surface hoar

15

is more common in sheltered areas below tree-line than above, further field work might

find suitable areas above tree-line for skiing on subsequent days.

Finally, forecasters might identify the conditions that would cause them to re-assess

the forecast. For example, a forecaster might ask the technicians to report back if natural

avalanches or substantial drifting are observed.

In this example, the forecaster's request for additional data, especially snowpack and

stability tests in and near avalanche start zones illustrates the fact that forecasts are

iterative and are constantly being refined (LaChapelle, 1980). Also, snow stability—even

in this simplified example—cannot be separated from terrain parameters such as the

direction a slope faces (aspect) or elevation band (above or below tree-line). Some

observations such as the “moderate” shear on surface hoar help the forecaster build an

intuitive, deterministic model; others such as wind direction pertain to the spatial

distribution of snowpack properties (Buser and others, 1985).

1.8 Computer Assisted Forecasting

Avalanche forecasting models may be either based on rules developed by experts

(e.g. Giraud, 1993; Schweizer and Föhn, 1995) or based on data. Data-based models

include discriminant analysis (Judson and Erikson, 1973; Bovis, 1977; Obled and Good,

1980), time series analysis (Salway, 1976), nearest-neighbours (Buser, 1983; Buser and

others, 1985; Buser, 1989; McClung and Tweedy, 1994; Kristensen and Larsson, 1995;

Blattenberger and Fowles, 1995), ordinary and logistic regression (Blattenberger and

Fowles, 1995), neural networks (Schweizer and others, 1994; Stephens and others, 1995)

or classification trees (Davis and others, 1993; Boyne and Williams, 1993; Davis and

Elder, 1994, 1995).

Nearest neighbour models compare recent values of a set of meteorological and

snowpack variables with the values of the variables on past days within the data base. The

avalanche activity on days with similar conditions (nearest neighbours) can be used to

estimate a probability of avalanching. However, in operational testing, forecasters show

less interest in the probability and more interest in the avalanche activity on days with

16

similar conditions (Buser, 1983). Specifically, a nearest neighbour model might identify

that present conditions are similar to some avalanche conditions that occurred prior to the

forecaster’s employment in the area. By assessing avalanche activity on similar days,

forecasters can anticipate avalanche activity by size, aspect and elevation using their

experience. Nearest neighbour models have been successfully tested in forecasting

operations in Switzerland (Buser, 1989) and Canada (McClung and Tweedy, 1994).

A related non-parametric method, the classification tree algorithm, uses past data to

build a hierarchy (or tree) of two-way decisions involving the forecasting factors. The

terminal nodes (or leaves) of the tree are the predicted levels of avalanche activity.

Classification trees are used in Chapter 9.

Non-parametric methods work with a mix of quantitative and qualitative data, do

not require normalizing transformations of non-normal data, and are sensitive to

non-monotonic relations between the forecasting variables and avalanche activity variable.

In contrast to data-based models, knowledge-based models (expert systems) do not

require an extensive data set. After interviewing experts, rules are constructed to reflect

their logic (McClung, 1995). Such systems can use both qualitative data such as the

character of the snow surface prior to the recent storm and quantitative data such as

meteorological variables. A recent expert system (Schweizer and Föhn, 1995), correctly

estimated the level of avalanche danger on a 7-level scale in the Alps near Davos on 73%

of days during the three winters, and was within one level of the verified avalanche danger

on 98% of days.

This dissertation focuses not on forecasting models but on stability indices for both

terrain-selection decisions and mesoscale forecasts. Such indices can be incorporated into

data-based models once there are sufficient data, or into knowledge-based models.

1.9 Atypical Snowpack Characteristics of Accident Avalanches

Based on reports of fatal avalanches in Canada between 1972 and 1991, Jamieson

and Johnston (1992a) estimate that 99% were slab avalanches, 87% were dry, and 93%

were triggered by people, mostly skiers. These characteristics are reflected in the present

17

study which is restricted to dry slab avalanches and emphasizes skier-triggered avalanches

over naturally occurring avalanches.

Although most slab avalanches start in weak layers consisting of precipitation

particles or partly decomposed precipitation particles, most fatal avalanches start in weak

layers with persistent forms such as surface hoar (Figure 1.11), faceted crystals and depth

hoar. Such crystals are slow to change shape and gain strength when subjected to

rounding metamorphism. As a consequence of their persistence, they remain weak when

deeply buried by subsequent snowfalls, and their failures result in thick—often large—slab

avalanches.

1.10 Skier-Triggering of Persistent Weak Layers

In a laboratory study of shear strength under various constant strain rates, the

ductile-brittle transition for manufactured depth hoar was between 8 x 10-5 and 2 x 10-4 s-1

(Fukuzawa and Narita, 1993). In a similar study of rounded grains, the ductile-brittle

transition was 4 x 10-4 to 8 x 10-4 s-1 (Narita and others, 1992). This implies that rounded

grains can be sheared 4 to 5 times faster than depth hoar before brittle fracture. Since

18

Figure 1.11 Microstructure of failure plane for fatal slab avalanche accidents in Canada,1972-91 (Jamieson and Johnston, 1992a). The pie charts are based on 34 of 45accidents with amateur decision-makers and 16 of 17 with professional decision-makersfor which the failure plane was reported.

skiers are believed to directly trigger brittle failure (Schweizer and others, 1995), depth

hoar which exhibits brittle failure for a wider range of strain rates is more sensitive to

skier-triggering than rounded grains. Although there are no published laboratory or field

studies for shear tests of surface hoar, a ductile-brittle transition similar to that of depth

hoar and consequent sensitivity to skier-triggering are expected.

1.11 Snow Profiles and Snowpack Tests

This section introduces the snow profile and four snow pack tests commonly used in

Canada and elsewhere. The last two tests described, namely, the shear frame test and the

rutschblock test, are the focus of this dissertation

1.11.1 Snow Profile

The snow profile is a systematic observation of snowpack layers (CAA, 1995) made

in a pit dug where the snowpack was undisturbed. Identification of weak layers is a

primary objective of snow profiles. The information recorded for each layer commonly

includes grain type and size (microstructure), resistance to penetration, liquid water

content and density, along

with a temperature profile.

“Hand hardness” is a simple

and widely used measure of

resistance to penetration

(Figure 1.12, Appendix A).

Interpretation of snow

profiles requires training and

experience. Mechanical tests

such as the shovel test

(Figure 1.13) and

compression test

(Figure 1.14) are often the

final stage of a profile. See

19

Figure 1.12 Testing hand hardness during a snow profileobservation. Ruler is used to identify position and thicknessof layers. Thermometers are used to measure snow surfacetemperature in shade and a temperature profile through thesnowpack. In this photo, snowpack layers were revealed bybrushing. (M. Shubin photo)

CAA (1995) for a detailed descriptions of

these tests.

1.11.2 Shovel Shear Test

For the shovel test, slope-parallel manual

force is applied to a shovel placed behind a

column of snow and progressively increased to

apply shear stress to the weak layers (Figure

1.13). Failures more than 0.2-0.3 m below the

bottom of the shovel blade are more likely due

to bending rather than to shear (Schaerer,

1989, 1991; CAA, 1995). Although the force

required to cause planar failures 0.2 m or less

below the bottom of the shovel are rated “very

easy”, “easy”, “moderate” or “hard”, the test

is used primarily to identify weak snowpack

layers rather than to quantify their strength.

The force rating cannot be directly related to

avalanching since it does not consider the

downslope shear stress due to the weight of

the slab on slopes or due to other triggers

such as skiers.

1.11.3 Compression Test

For the compression test (Figure 1.14),

a sequence of vertical blows of increasing force

is applied by the hand to a shovel blade placed

20

Figure 1.13 The shovel shear test usedprimarily to identify weak snowpacklayers. (J. Hughes photo)

Figure 1.14 Compression test. Failuresare visible on the smooth walls of thecolumn. (J. Hughes photo)

on top of a column of snow and the force required to cause visible failure is rated “very

easy”, “easy”, “moderate” or “hard”. Experience in the Canadian Rocky Mountains

suggests that increased tapping force correlates with decreased slab avalanching due to

triggers such as skiers or explosives (CAA, 1995).

1.11.4 Shear Frame Test

Various shear frame tests have been used in Switzerland and Canada since the

1960’s (Roch, 1966a, b; Schleiss and Schleiss, 1970). After the weak layer to be tested

has been identified in a profile or by another snowpack test, overlying snow is carefully

removed to within 15-50 mm of the weak layer (Figure 1.15). A sheet metal frame, usually

slightly trapezoidal and usually with two or more intermediate cross-members is placed in

the snow above the weak layer so

that the lower edge of the frame is

close to—typically 0-10 mm

above—the weak layer to be

tested. Commonly, the operator

slides a blade around the frame to

make sure that the frame is not

adhering to the surrounding snow.

A force gauge that records the

maximum force is attached to a

hook or cord on the front of the

frame and manual force is applied

to cause a shear failure in the weak

layer. Details of the technique are

described in Section 3.3. The shear

strength, obtained by dividing the

maximum force recorded by the

21

Figure 1.15 Shear frame test. (J. Hughes photo)

gauge by the area of the frame, is used in various formulas for stability indices which are

summarized in Section 2.4.

1.11.5 Rutschblock Test

The rutschblock (“glide” block)

test is a slope stability test first used by

the Swiss army to find weak snowpack

layers (Föhn, 1987b). On an undisturbed

slope preferably inclined at 25° or

steeper, a column of snow 2 m wide

(across the slope) and 1.5 m down-slope

is isolated from the surrounding

snowpack by shovelling or cutting with a

cord, saw or tail of a ski (Figure 1.16).

The block is progressively loaded in

seven steps by a skier as described in

Section 3.6. The rutschblock score is

simply the loading step (1-7) at which a

weak layer in the column fails, allowing

the upper portion of the column (the

block) to displace downslope. Further

details on the rutschblock technique and

method of scoring are described in Section 3.3.

1.12 Objective and Outline

The objectives of this dissertation are:

to improve shear frame stability indices, and

to investigate the merit of shear frame stability indices and rutschblock scores for

assessing the stability of slabs overlying persistent weak layers which cause most

fatal avalanches in Canada.

22

Figure 1.16 Rutschblock test showingdisplaced block. (M. Shubin photo)

Although naturally occurring slabs are considered, the emphasis is on skier-triggered slabs,

which are the primary concern in backcountry skiing.

Chapter 2 reviews relevant literature on the rutschblock and shear frame tests and on

shear frame stability indices. Chapter 3 describes the shear frame and rutschblock testing

techniques. Field and finite element studies of the shear frame test are summarized in

Chapters 4 and 5, respectively. Chapter 6 summarizes the results of field studies relating

shear frame stability indices from avalanche start zones and safe study sites to avalanche

activity. Similar field studies relating rutschblock results to avalanche activity are

presented in Chapter 7. Case studies of 5 skier-triggered avalanches at which the shear

frame stability index for skier triggering and/or the rutschblock test incorrectly indicated

stability are presented in Chapter 8, and used to draw conclusions about the limitations of

stability tests and about primary failures on shallow slopes. Chapter 9 develops

multivariate forecasting models based on classification trees to assess whether shear frame

stability indices and rutschblock scores can improve forecasting based on meteorological

variables alone. The conclusions and recommendations are presented in Chapters 10 and

11, respectively.

23

24

2 LITERATURE REVIEW

2.1 Introduction

This dissertation focuses on shear frame stability indices and rutschblock tests for

assessing the stability of dry slabs overlying persistent weak snowpack layers, with an

emphasis on skier-triggered slabs. Before reviewing research on these stability tests, slab

failure models are summarized in Section 2.2 and the effects of shear frame size, design

and normal load on shear strength are reviewed in Section 2.3.

Two types of shear frame stability indices are considered:

slope-specific stability indices obtained using shear frame tests on avalanche slopes

which assess the stability of the tested slope (Section 2.4), and

extrapolated stability indices obtained at safe study sites (level or sloping) which

assess avalanche activity in surrounding terrain (Section 2.5).

While the site-specific stability indices have been used by researchers to assess the

predictive value of the indices and spatial variability of stability within avalanche start

zones (Section 2.6), extrapolated stability indices are used for operational forecasting and

are undergoing refinement by researchers.

Research on the rutschblock test technique and assessments of correlations with

avalanche activity are reviewed in Section 2.7.

2.2 Slab Failure

Before a slab can release as an avalanche, fractures must occur in the weak layer at

the base of the slab and around the slab at the crown, flanks and stauchwall (Figure 2.1).

Bucher (1948) proposed that the primary failure that initiates avalanche release could start

in four different ways as summarized in Table 2.1. Failure could occur in any of these

locations whenever stress exceeds strength.

Until 1970, opinions of researchers varied regarding which of the four cases in

Table 2.1 was most common, and consequently most important. Haefeli (1963, 1967)

emphasized crown tension fracture (Case 3) but did not verify his hypothesis with field

25

data. Bradley (1966) and Bradley and Bowles (1967) focused on compressive failures of

thick layers of depth hoar (Case 2). Bradley and Bowles (1967) provided limited field data

to support a correlation between the ratio of resistance-to-vertical-penetration to vertical

stress due to slab weight and avalanching initiated by the collapse of thick depth hoar

layers in a Continental snowpack, similar to that of the Canadian Rocky Mountains. Roch

(1966a) emphasized basal shear failure (Case 1) and supporting field data are discussed in

Section 2.3. In a

variation of Case 1,

Perla and LaChapelle

(1970) argued that the

first fracture was in

tension at the crown but

that the tensile stress at

the crown was caused

by a loss of shear

support (ductile failure)

in the weak layer. This

emphasis on ductile

shear failure was

supported by

26

Table 2.1 Possibilities for Primary Fracture (Bucher, 1948)

Case Location Fracture Associated Conditions

1 central area of slab(neutral zone)

shear in weak layer “loose” weak layers or surfacehoar

2 central area of slab(neutral zone)

compression at base ofsnowpack

depth hoar at base of snowpack

3 crown region(slab boundary)

tension through slab fresh storm snow

4 flank region(slab boundary)

shear through slab infrequent

Figure 2.1 Perspective diagram showing slab nomenclature,orthogonal axes for x, y and z co-ordinates, slope inclination fromthe horizontal, Ψ, and slab thickness, h, measured vertically.

McClung’s (1977, 1979) laboratory studies of snow under slow shear deformation. Recent

slab failure models (McClung, 1981, 1987; Bader and Salm, 1990) have focused on shear

failure and propagation within the weak layer at the base of the slab (Case 1). Although

research has not proven that primary tensile fracture at the crown (Case 3) and primary

shear fracture at a flank (Case 4) do not occur, such methods of initiating slab failure lack

supporting field data and have not attracted recent interest.

The primary compressive failures reported by Bradley and Bowles (1967) and field

workers in areas with depth hoar layers may not be in conflict with the model of shear

failure and propagation within weak layers developed by Bader and Salm (1990). Their

model requires high shear stress concentrations and high shear strain rates which are more

likely in thin weak layers and at interfaces than in thick weak layers associated with

primary compressive failures by Bradley (1966) and Bradley and Bowles (1967). Further,

reports of substrata consisting of relatively weak depth hoar (that did not release)

extending to the crown fracture (T. Auger, personal communication) suggest that,

regardless of the character of the primary failure, the subsequent fracture propagates along

the interface. Current research on slab failure and field studies of stability indices,

including this dissertation, are based on primary shear failure of the weak layer (Case 1).

Primary compressive failures are revisited in Chapters 7 and 8, for areas with thick layers

of depth hoar, and on shallow slopes where compressive stress exceeds shear stress.

2.3 Shear Frame

Shear frames have become the most common device for testing the shear strength

of weak snowpack weak layers. An alternative, the rotary shear vane similar to that used

for testing soils has been used to test homogeneous snow layers (Keeler and Weeks, 1967;

Martinelli, 1971; Perla and others, 1982; Brun and Rey, 1987). The shear vane test is

faster than the shear frame test since it can be pushed into the snow with a shaft and hence

does not require removal of most of the snow above the test layer. However, since the

bottom of the vanes cannot be consistently positioned in or near thin weak layers, it was

deemed unsuitable for the present study.

27

2.3.1 Shear Frame Size

For many materials, mean strength decreases with the cross-sectional area or

volume of the specimen since larger specimens can contain larger flaws (Griffith, 1920).

For shear frame tests, the area of the fracture surface is equal to the area of the frame, and

consequently a decrease in mean strength is expected with an increase in frame area.

Roch (1966a, 1966b) introduced the shear frame for measuring the shear strength

of weak snow layers. He used a frame with an area of 0.01 m2.

Perla (1977) compared mean strengths from frames with areas of 0.01, 0.025,

0.05, 0.10 and 0.25 m2. In each of 7 comparisons based on 100 paired tests with frames of

different areas, the larger frame gave a lower mean strength than the smaller frame.

Stethem and Tweedy (1981) also found that a larger frame (0.025 m2) resulted in a lower

mean strength compared to a smaller frame (0.01 m2). In four of five comparisons with

frames ranging from 0.01 to 0.25 m2, Föhn (1987a) found mean strengths for the larger

frame were lower than for the smaller frame.

Sommerfeld (1973), Sommerfeld and others (1976), Sommerfeld and King (1979),

and Sommerfeld (1980) proposed that size effects in shear frame tests could be explained

by Daniels’ (1945) thread bundle statistics. Using this theory, Föhn (1987a) compiled his

results with those of Perla (1977) and Sommerfeld (1980) to obtain a curve of correction

factors. For frames larger than 0.3 m2, mean shear strengths asymptotically approached the

strength of an arbitrarily large specimen. This asymptote, called the Daniels strength, can

be obtained by multiplying the mean strength obtained with a particular area of frame by

the appropriate correction factor. For frames with areas of 0.01, 0.025 and 0.05 m2, the

correction factors are 0.56, 0.65 and 0.71 respectively (Sommerfeld, 1980; Föhn, 1987a).

Although very large frames may seem advantageous, frames larger than 0.1 m2 are

less practical because:

the necessary manual pull forces cannot be applied consistently and smoothly by a

single operator,

curvature of thin weak layers which is common in avalanche starting zones makes

aligning large frames more difficult, and

28

for shear frame tests on steep slopes, adjustments to the shear strength for the

slope-parallel force due to the weight of the frame and snow in the frame would

become increasingly important.

Three sizes of smaller frames (0.01, 0.025 and 0.05 m2) are in use. Perla and Beck

(1983), Sommerfeld (1984) and Jamieson and Johnston (1993a) prefer a 0.025 m2 frame.

Föhn (1987a) prefers a 0.05 m2 frame and Schaerer (1991) prefers the 0.01 m2 frame used

at Rogers Pass and Kootenay Pass.

2.3.2 Shear Frame Design

To distribute the applied stress more evenly through the snow layer being tested,

Roch’s (1966a) frame had two intermediate cross-members. The relatively rigid outer

frame distributes the manually applied load onto the rear cross-member and the two

intermediate cross-members. The lower tip of each of the active surfaces (cross-members)

creates a shear stress concentration in the snow layer.

Perla and Beck (1983) stated that the stress concentrations are influenced by the

ratio of the height of the cross-member, d, to the length of the snow sub-specimen in front

of the cross-member, w (Figure 2.2). In a

field comparison, they decreased d/w from

0.75 to 0.37 by increasing the number of

intermediate cross-members from 2 to 5 and

found that mean strength measurement was

reduced by 15%. Further, they stated that

increasing d/w should increase the normal

load (and hence increase the strength) and

might contribute to disturbance of the weak

layer when the frame is inserted (Perla and

Beck, 1983).

Roch’s (1966a) 0 .01 m2 frame had 3

active cross-members and a d/w ratio of

0.75. Perla and Beck (1983) and

29

Figure 2.2 Shear frame showing rearcross-member and two intermediatecross-members that distribute the load.

Sommerfeld (1984) retained the slightly trapezoidal shape and three active cross-members

but preferred frames with an area of 0.025 m2. Perla and Beck (1983) maintained the d/w

ratio of 0.75 whereas Sommerfeld (1984) reduced the frame height to obtain a d/w ratio of

0.4.

In contrast to the shear frames with compartments, Brown and Oakberg designed a

frame which distributed the load more evenly using thirty-two 10 mm wide fins anchored

to a plate on top of the 0.01 m2 shear frame shown in Figure 2.3 (Lang and others, 1985).

The fins were approximately

10 mm apart and extended

17 mm down into the 25 mm

high frame. Lang and others

(1985) used the frame for five

sets of 10 tests on a particular

surface hoar layer that ranged in

strength from 0.03 to 0.35 kPa

over a 40 day period, but

reported that the surface hoar

layer was sometimes too weak

to test. Maximum variability

was reported to be less than

0.20 kPa for 10 tests on the

same day. No comparison of this frame with conventional compartmental frames was

found in the literature.

Föhn (1987a) used a 0.05 m2 frame with cross-members and a d/w ratio of 0.64.

This “Swiss” frame is constructed of 1.5 mm stainless steel and is four times as heavy as

the 0.025 m2 frame adopted as a standard for the present study. Using compartmental

frames, Perla and Beck (1983) report an increase in strength with frame weight. No

comparison of the Swiss frame with other frames was found in the literature.

30

Figure 2.3 Underside view of finger-fin frame. Load isdistributed onto 32 finger-fins, each extending 17 mminto the 25 mm high frame.

Field comparisons between various frames, including those that varied in

cross-member height and spacing between cross-members, are presented in Chapter 4.

Finite element studies of shear stress for various frame designs are presented in Chapter 5.

2.3.3 Effect of Normal Load on Shear Strength

The shear strength of granular materials generally increases with normal load. For

failures due to yielding, a linear increase with normal load is commonly modelled with the

Mohr-Coloumb failure criteria (Holtz and Kovacs, 1981, p. 453.). However, since most

shear frame tests are pulled fast enough to cause brittle fractures (Section 4.5), a linear

Mohr-Coloumb effect should not be assumed (de Montmollin, 1982).

To assess the effect of normal load on shear strength, Roch (1966b) experimented

with weights placed on top of shear frames. He measured an increase in shear strength

with an increase in the normal load. Perla and Beck (1983) also reported an increase in

shear strength with frame weight but questioned whether this normal load effect was due

to “internal friction”, an inertial effect associated with the rapid pull on the frame, or

ploughing of the weighted frame in the substratum. Assuming the normal load effect was

due to internal friction, φ, Roch (1966a, b) expressed the adjusted shear strength as

Σφ = Σ + σZZ φ (2.1)

where the shear strength, Σ, is the maximum pull force divided by the area of the shear

frame and the normal stress on the weak layer due to the slab of density, ρ, slab thickness,

h (measured vertically) on a slope of inclination, Ψ, is

σZZ = ρgh cos2Ψ (2.2)

Roch (1966b) found that the internal friction term φ depended on strength and

microstructure. Using a 0.01 m2 frame to obtain a strength, Σ100, he determined empirical

formulas for φ for several different microstructures

fresh snow φ(Σ100, σΖΖ) = 0.1 + 0.08 Σ100 + 0.04 σZZ (2.3a)

rounded grains φ(Σ100, σΖΖ) = 0.4 + 0.08 Σ100 (2.3b)

facets and depth hoar φ(Σ100, σΖΖ) = 0.8 + 0.08 Σ100 - 0.01 σZZ (2.3c).

31

Perla and Beck (1983) argued that the normal adjustment was not “crucial” since

the coefficient for the correlation between their unadjusted stability index (Section 2.5)

and the normal load σZZ was only r = -0.44 for 23 slab avalanches. However, the

significance of this correlation is p < 0.04 indicating that the correction for normal load

may have merit.

Field data and analysis of normal load effects from the present study are presented

in Section 4.11.

2.4 Slope-Specific Stability Indices

Field studies of stability indices obtained from shear frame tests at avalanche start

zones and assessed using the avalanche activity of those start zones are reviewed in this

section. The indices vary depending on whether they include stress due to artificial triggers

in the denominator, and adjust for normal load due to the slab, or the size of the frames

(Table 2.2).

Assuming that most slab failures start with shear failure of the weak layer, Roch

(1966a) began field studies of slab stability based on a stability index

(2.4)SRoch =Σ100 + σzz φ(Σ100, σzz)

σxz

where σXZ is the shear stress in the weak layer due to the weight of the overlying slab.

From statics, the shear stress due to the slab (Figure 2.1) is

σXZ = ρgh sin Ψ cos Ψ (2.5)

Near 35 avalanches, Roch (1966a) found that the index SRoch ranged from 0.76 to

7.5, averaged 2.05 and had a standard deviation of 1.20. (He believed that when slopes

avalanched with SRoch > 2, the primary fractures must have been tensile fractures at the

crown—exceptions to the more common shear failures within weak layers.) However,

Roch did not compare values of SRoch from slopes that had avalanched with slopes that had

not. Also, although he reported the trigger for each of the avalanches, he did not separate

the 24 naturally triggered avalanches from the 11 artificially triggered avalanches in his

analysis, even though the denominator of the strength/stress index (Eq. 2.4) did not

include the superimposed stress of artificial triggers such as skiers or explosives.

32

Using a 0.025 m2 shear frame to obtain a measure of shear strength, Σ250, Perla

(1977) used the ratio Σ250/σXZ (without the normal load correction) for field studies of slab

stability. For 80 slab avalanches, the ratio ranged from 0.19 to 6.4, averaged 1.66 and had

a standard deviation of 0.98. He believed that the large standard deviation cast doubt on

the usefulness of the ratio for stability evaluation. However, Perla did not compare values

33

Table 2.2 Field Studies of Shear Frame Stability Indices

Study TrueIndex

Adjust forSize Effects

Adjust forNormal Load

Include ArtificialStress

Slope-Specific Stability Indices

Roch (1966a, b) Y N Y3 N

Perla (1977) Y N2 N2 N

Sommerfeld and King (1979) Y Y N N

Conway and Abrahamson(1984)

Y N Y N

Föhn (1987a) Y Y Y3 Y

Conway and Abrahamson(1988)

Y N Y Y

Föhn (1989) Y Y Y3 Y

Jamieson and Johnston(1995a)

Y Y Y3 Y

Extrapolated Stability Indices1

Schleiss and Schleiss (1970) N N N N

Stethem and Tweedy (1981) N N2 N N2

Jamieson and Johnston(1993)

Y Y Y3 natural av. only

Jamieson and Johnston(1995a)

Y Y Y3 Y

1 Study relates stability parameter from shear frame tests at safe study site to avalanche activity in surrounding terrain.2 Effect studied but not incorporated into stability index/ratio.3 Applied normal load adjustment for granular snow to weak layers with various microstructures.

of Σ250/σXZ on slopes that had avalanched with values on slopes that had not, and did not

report the type of trigger for the avalanches.

On eight slopes that had recently avalanched, Sommerfeld and King (1979)

measured the shear strength of the failure plane with a 0.025 m2 frame. They adjusted the

shear strength (without the normal load correction) Σ250 for size effects to obtain the

Daniels strength of an arbitrarily large specimen, Σ∞. Remarkably, in five of the eight

cases, the ratio Σ∞/σXZ was between 0.97 and 1.04. However, in the remaining three cases,

the ratios were 0.83, 1.68 and 2.47. Further, they only tested eight slopes, and did not

include the stress due to skiers or explosives in the stress term (denominator of stability

index) although two of the slopes were triggered by skiers and three were triggered by

explosives.

Conway and Abrahamson (1984) used a different technique for their shear frame

tests. Working on avalanche slopes, they isolated a vertical column of snow and embedded

the shear frame on top of the column, which in some cases extended over 1 m above the

weak layer. While this technique includes the inherent effect of normal load on the weak

layer and is independent of shear frame size, pulling the frame downslope superimposes

substantial bending stress on the shear stress in the weak layer. Eight slabs that avalanched

yielded stability indices that averaged 1.57 with a standard deviation of 1.29 in contrast to

18 slabs that did not avalanche where the stability index averaged 4.25 with a standard

deviation of 2.78. This is the first comparison of stability indices on slopes that did and did

not avalanche, and the results indicate the merit of stability indices for discriminating

between stable and unstable slopes, despite the effect that bending may have had on the

results.

Föhn (1987a) combined Roch’s (1966a, b) normal load adjustment with

Sommerfeld and King’s (1979) size correction to obtain a stability index for natural

avalanches

(2.6)S =Σ∞ + σzzφ

σxz

and added the term ∆σXZ for the artificially induced stress into the denominator of

Equation 2.6 to obtain an index for artificially triggered avalanches

34

(2.7)S =Σ∞ + σzzφσxz + ∆σxz

By assuming isotropy and linear elastic behaviour and ignoring deviatoric stress gradients,

Föhn derived formulas for estimating ∆σXZ for a walker, a skier, a “snowcat” and a 1 kg

explosive. For a skier on a slab of thickness h on a slope of inclination Ψ, ∆σXZ is

(2.8)∆σxz =2L cos αmaxsin αmax

2 sin(αmax + Ψ)πh cosΨ

where L is the line load due to a skier (500 N/m) and αmax is the angle from the snow

surface to the peak shear stress (Figure 2.4) tabulated by Föhn (1987a) for common values

of Ψ. For a skier on a 38° slope, ∆σXZ simplifies to 0.14/h kPa where h is in m. Föhn’s

(1987a) formula for ∆σXZ, the slope-parallel static stress induced by a skier, was verified

by Schweizer’s (1993) finite element model and approximately by field measurements.

Using a load cell buried at various depths between 0.1 and 0.6 m under a level snow

surface, Schweizer and others (1995) showed that calculated normal stress due to a static

skier agreed well with the normal stress generated by skiers pushing vertically downwards

with their legs. Also, Schweizer and others (1994) showed that the stress on the load cell

depended on the properties of the

slab, a factor that is ignored in

analytical formulas for skier induced

stress used by Föhn (1987a) and

Jamieson and Johnston (1995a).

Using S for natural

avalanches and S' for slabs triggered

mainly by skiers or explosions, Föhn

rated the combined “success” of S

and S' for discriminating between

snow slopes that had, and had not

avalanched. S or S' is rated

successful where the index is less

than 1 and the slab released, or is

35

Figure 2.4 Cross-section of slab showing locationof peak shear stress induced by static skier.

greater than 1.5 where the slab did not release. Values of S or S' between 1 and 1.5 were

considered to indicate transitional stability and were excluded from the success score. The

success score for S combined with S' for 110 avalanche slopes was 75% (Föhn 1987a).

Since avalanche forecasting typically relies on at least 10 factors (Section 1.7), a single

variable capable of predicting the stability of 75% of avalanche slopes is promising. Föhn

(1987a) along with Roch (1966a) and Jamieson and Johnston (1993a, 1995a) applied

Roch’s normal load correction for granular snow (Eq. 2.3b) to all weak layers,

independent of their microstructure.

To account for the fact that skis penetrate into soft snow, often by as much as

0.3-0.5 m, thereby decreasing the distance from the skis to the weak layer and increasing

the skier-induced stress, Jamieson and Johnston (1995a) adjusted Föhn’s (1987a)

skier-stability index S' to allow for ski penetration to obtain SK and related it to avalanche

activity on persistent weak layers. Based on an additional winter of field data since

submission of Jamieson and Johnston (1995a), this adjustment is refined in Chapter 6 and

the refined index is assessed using a larger set of avalanche data.

While the ability of slope-specific stability indices to discriminate between stable

and unstable slopes has been established (Conway and Abrahamson, 1984; Föhn, 1987a,

Jamieson and Johnston, 1995a), most avalanche forecasting and control programs do not

have the resources to do shear frame tests in more than 1 or 2 avalanche start zones—and

it is not always safe to do so. Consequently, the application of shear frame results to such

operations requires that stability indices be extrapolated to surrounding terrain from shear

frame tests at sites that are generally safe to access.

2.5 Extrapolated Stability Indices

While a slope-specific stability index depends on the slope inclination of the start

zone, an extrapolated stability index intended as a predictor of avalanche activity in

surrounding terrain should apply to start zones with various slope inclinations. Either the

index can be calculated for a minimum or average inclination, or the trigonometric

functions of slope inclination can be dropped, resulting in an expression of the form

36

Σ/ρgh—which is the ratio of slope-parallel shear strength to vertical stress due to the

weight of the slab. Historically, this ratio has been called the stability factor (Schleiss and

Schleiss, 1970; Salway, 1976; Stethem and Tweedy, 1981; NRCC/CAA, 1989; Jamieson

and Johnston, 1993a). However, since Σ/ρgh is not a stability factor or stability index as

defined in some engineering texts, the Canadian Avalanche Association now refers to

Σ/ρgh as the Stability Ratio (CAA, 1995). Jamieson and Johnston (1993a) showed that

when stability indices such as Föhn’s S are calculated for a constant slope angle, they are

approximately proportional to ratios of the form Σ/ρgh. Consequently, ratios of the form

Σ/ρgh are included as shear frame stability indices in this study.

Since 1963, the avalanche control program for the Trans-Canada Highway through

Rogers Pass has used ratio Σ100/ρgh, where Σ100 is the shear strength measured with a

0.01 m2 frame, as an index of stability for avalanche paths that can affect the highway (D.

Skjönsberg, personal communication). Schleiss and Schleiss (1970) report that snow

stability in nearby start zones is critical when the ratio Σ100/ρgh (measured in a level study

plot) is less than 1.5. Further south in the Selkirk Mountains, the avalanche control

program for the highway through Kootenay Pass uses the same ratio and critical level (J.

Tweedy, personal communication).

Stethem and Tweedy (1981) report Σ100/ρgh and Σ250/ρgh values of 0.97 and 1.02,

respectively, measured in a level study plot near the time that natural avalanches released,

and 1.87 and 1.29, respectively, when avalanches were artificially triggered. The higher

values for artificially triggered slabs supports the use of a term for artificially induced

stress in the denominator of stability indices, the magnitude of which will depend on the

type of trigger (skier, explosive, etc.).

To extrapolate from representative study slopes to surrounding avalanche slopes,

Jamieson and Johnston (1993a) calculated S35 which is simply Föhn’s S (Eq. 2.6)

calculated for a 35o slope, an inclination typical of many avalanche starting zones. Based

on 70 test days over three winters in the Cariboo and Monashee Mountains (Fig. 2.5),

Jamieson and Johnston chose the critical value of S35 empirically, weighting days with dry

slab avalanches more than days without such avalanches. Transitional stability was defined

37

as the band of S35 values within ±10% of the critical value. S35 scored a success when one

or more dry natural slab avalanches were reported with 15 km of the study site and S35

was below the transition band, or no dry slab avalanches were reported and S35 was above

the transition band. Based on this criterion, S35 correctly predicted avalanche activity on

75-87% of the 70 test days, depending on whether S35 was measured at a level plot or an

inclined study slope, and whether natural dry slab avalanches with estimated dates were

included or excluded. For S35 measured in a level study plot, the avalanche activity,

including avalanches with estimated dates, is shown in Figure 2.5.

Most of the failure planes for the avalanches in this study were within the more

recent storm snow. S35 has not been assessed for the deeper, more persistent weak layers

typical of most fatal avalanches. Also, Jamieson and Johnston applied Roch’s normal load

correction for granular snow (Eq. 2.3b) to all weak layers, independent of their

microstructure.

Jamieson and Johnston (1995a) used Ψ = 35° in the equation for the slope specific

equation, SK, (Section 2.4) to obtain an extrapolated index for skier stability, SK35. A

38

Figure 2.5 Avalanche activity and concurrent values of S35 from Caribooand Monashee Mountains, 1990-92. (After Jamieson and Johnston,1993a) Only class 1.5 and larger natural avalanches (CAA, 1995) areincluded since smaller natural avalanches are not reported consistently,and do not pose a serious threat to skiers.

refinement denoted by SK38 which differs from SK35 in slope inclination and normal load

adjustment based on recent studies in the Columbia Mountains (Section 4.11), is assessed

using a larger data set than previously available in Chapter 6.

2.6 Flaws in Weak Layers and Spatial Variability of Stability Indices

Snowpack properties, including the mechanical properties of weak layers, vary

within avalanche start zones. For natural avalanches that release under their own weight,

Bader and Salm (1990) argued that the critical shear strain rate for propagation could only

be achieved at flaws in weak layers with insufficient shear strength to resist the shear

stress of the overlying slabs. Such flaws have been called shear bands, slip bands, or slip

surfaces (Palmer and Rice, 1973; Rice, 1973, Singh, 1980; McClung, 1981; McClung,

1987), deficit zones (Conway and Abrahamson, 1984) and super-weak zones (Bader and

Salm, 1990). Most of the theoretical models of snow-slab failure (Singh, 1980; McClung

1981; McClung, 1987) are based on fracture mechanical models for clay slab failures

(Palmer and Rice, 1973; Rice, 1973). According to the analytical and finite element model

of Bader and Salm (1990), super-weak (deficit) zones can grow over periods of up to 60

minutes prior to reaching a critical size (5-30 metres in downslope length) beyond which

brittle fractures release the slab. If super-weak zones can exist for such periods of time,

then field tests of snow stability may, potentially, verify the theory.

Based on shear frame tests spaced along a 15 m wide part of a crown, Conway

and Abrahamson (1984) report that their stability index varied from < 1 to 3.6. Also, for

shear frame tests down a 50 m long portion of a flank at a second avalanche, the index

varied from < 1 to 2.3. Shear frame tests for which the weak layer fractured before the

pull could be applied were assigned a stability index < 1 whereas Föhn (1989) (and

presumably other authors) rejected tests involving “pre-fractures”. Conway and

Abrahamson used stability indices with values < 1 as evidence of super-weak (deficit)

zones, although such results only occurred where the weak layer fractured before the pull

was applied to the frame (Föhn, 1989).

39

In a subsequent study of 5 slopes, Conway and Abrahamson (1988) used

Vanmarke statistics to determine a probability distribution for the super-weak areas of five

slopes, four of which avalanched and one which fractured locally. Since each of the slopes

had a 95% probability of including a deficit zone at least 2.9 m long, they concluded that

small deficit zones could determine the stability of avalanche slopes. Föhn (1989)

countered this idea by pointing out that

greater variability for stability indices—due to the presence of super-weak

zones—than for other snowpack properties has not been detected,

neither his nor Conway and Abrahamson’s (1984, 1988) data provide evidence of

deficit areas for natural avalanches (S < 1) if specimens that fractured before the

frame was pulled are rejected,

small deficit zones for artificial triggers (S′ < 1) exist on slopes that could not be

triggered, and

stability tests at a single point in an avalanche start zone have proven to be useful

indications of stability and this would not be true if small deficit zones determine the

stability of avalanche slopes.

Also, Jamieson and Johnston (1995a) report that 17 of 23 (83%) slopes on which

SK < 1 were skier triggered, and 14 of 17 (72%) slopes on which SK > 1.5 were not

skier-triggered. It is unlikely that stability tests from a single pit (area 2 m2) in an

avalanche start zone (area 100-20,000 m2) could have such predictive value if small deficit

zones determine stability. Additional field data to support the merit of tests for

skier-stability at a single location in an avalanche slope are presented in Chapters 6 and 7.

Also, the spatial variability of stability tests for skiers is documented for 9 slopes in

Chapter 7.

However, no field study of natural avalanches has disproven Bader and Salm’s

(1990) assertion that shear strain rates necessary for propagation can only be achieved at

the stress concentration around super-weak (deficit) zones. On the other hand, there is no

field evidence that super-weak zones can gradually grow in length to over 10 m during

periods of up to 60 minutes as calculated by Bader and Salm (1990). However, the fact

40

that skiers can produce dynamic

stresses in weak layers

comparable to the static stresses

due to the weight of the slab and

to the shear strength of weak

layers, indicates that super-weak

zones are not necessary for

skier-triggering (Schweizer and

others, 1995).

2.7 Rutschblock

The rutschblock test (Figure 2.6) is a slope stability test first used by the Swiss

army to find weak snowpack layers (Föhn, 1987b). Unlike the shear frame test, it does not

require extensive training or specialized equipment such as force gauges. The lower wall

of the column is exposed by digging a pit in the snowpack. The sides and upper wall can

be isolated from the surrounding snowpack by shovelling, cutting with cords, skis, or large

saws. The block is loaded in seven steps by a skier. The rutschblock score is simply the

loading step at which a weak layer in the column fails, allowing the upper portion of the

column (the block) to displace downslope. Further details on the rutschblock technique

are presented in Section 3.3.2.

Although the rutschblock test and variations of it have been promoted by earlier

publications in German (e.g. Munter, 1973), its popularity in North America follows

Föhn’s (1987b) paper in English. In a study involving rutschblock tests on 150 avalanche

slopes, Föhn compared each of the seven rutschblock scores with the relative frequency of

slab avalanches (Figure 2.7). The percentage of slab avalanches decreased with increasing

rutschblock score. However, since 10-15% of the slopes with the highest rutschblock

scores avalanched, the rutschblock test did not, by itself, indicate that a particular slope

was stable. This limitation of the rutschblock test is, according to Föhn, due to imprecise

site selection. Since rutschblock scores vary on any particular avalanche slope, one or two

41

Figure 2.6 Rutschblock test.

tests may miss the least stable

part of a slope. Föhn

emphasized that factors such as

snow profiles and weather must

be used along with rutschblock

tests to assess slab stability.

Whitmore and others

(1987) attempted to compare

the rutschblock test to other

“level-type tests” such as the

shovel shear test. In 16 cases,

individual rutschblocks were

compared with 1-4 level-type

tests. There were insufficient data

for conclusions since only 16 rutschblock tests were made and the rutschblock scores only

varied from 5 to 7. Nevertheless, the authors did develop a preference for the rutschblock

because of the variability of results of the

lever-type tests.

Munter (1991, p. 93-102) compared

the rutschblock test to a variation with a

wedge-shaped block called the rutschkeil

(Figure 2.8). When a cord is used to cut the

sides and upper wall, the rutschkeil test is

usually faster than the rutschblock test.

Munter also finds the rutschkeil “more

sensitive” than the rutschblock since the

block can be loaded more gently by a skier

moving onto the block from the side,

whereas the rutschblock requires a skier to

42

Figure 2.8 Rutschkeil test (after Munter,1991).

Figure 2.7 Percentage of slab avalanches andconcurrent rutschblock scores (after Föhn, 1987b).

step down onto the block from above. However, this difference seems unimportant since

both Munter (1991) and Föhn (1987a) considered the slope to be unstable whether the

block moves before stepping onto the block or while stepping onto the block.

One disadvantage of the rutschkeil test involves the variable area of the failure

surface. Experience with wedge-shaped blocks during the winters of 1989-90 to 1990-91,

shows that sometimes the whole wedge displaces indicating a shear failure over an area of

3 m2 and sometimes only the portion of the wedge downslope of the skier displaces

indicating a shear failure over an area of approximately 1.8 m2 and a fracture of unknown

character under the skis. (This problem is reduced for the rutschblock since the block is

loaded closer to the upper wall.)

Further, since the skis extend 0.1 to 0.2 m on both sides of the wedge, the skier

load is carried partly by the wedge being tested and partly by the surrounding snowpack.

Also, the portion of the skier load carried by the snow surrounding the wedge depends on

factors unrelated to stability, including the compressibility of the snow and the length and

stiffness of the skis. This problem and the variability of the area of the shear failure surface

make the rutschkeil test less suited to research on snow stability than the rutschblock test.

Munter (1991, p. 90)

reported that a 5° decrease in slope

inclination tends to increase the

rutschkeil score by approximately

one step, but did not provide

supporting data. Although there

are differences in the loading steps

for the rutschkeil and rutschblock

test, Jamieson and Johnston

(1993b) provided field data and a

simple numerical technique to

indicate that a 10° decrease in

slope inclination tends to

43

Figure 2.9 Cord-cut rutschblock (after Jamieson andJohnston, 1993b).

increase the rutschblock score by one step. However, this general effect was only

significant on 10 of 24 slopes, probably due to natural variability of snowpack properties

and the non-linear increase in stress with rutschblock loading steps reported by Schweizer

and others (1995).

Jamieson and Johnston (1993b) found that the time required to perform a

rutschblock could be reduced by approximately half by using a ski or specialized saw to

cut the sides and upper wall of the rutschblock. Provided no knife-hard crusts exist in the

slab to be cut, 4-6 mm knotted cords can also be used (Figure 2.9). Based on 21

comparisons between adjacent tests using these alternative block-cutting techniques,

Jamieson and Johnston were unable to detect any significant effect of the block-cutting

technique on the resulting rutschblock scores provided the width of the block flared as

shown in Figure 2.9.

In a study of 36-67 rutschblock scores on six uniform slopes that varied in slope

inclination by less than 10°, Jamieson and Johnston (1993b) found that one rutschblock

test has an approximately 67% probability of giving the median score for the slope and

97% probability of giving a score within 1 step of the slope median. The probability of the

median score of 2 independent tests being within ½ step of the slope median is

approximately 91%.

2.8 Summary

2.8.1 Slab Failure

Most cases of snow slab failure begin with a shear failure within the layer or weak

interface (e.g. McClung, 1987), although limited field data suggest that slab failure can

also start with a compressive (collapse) failure where thick weak layers exist in the

snowpack (Bradley and Bowles, 1967). For natural avalanches, shear failures start at

flaws in weak layers, grow in a ductile manner (Bader and Salm, 1990) and propagate

rapidly when they reach a downslope length of approximately 2-60 m. In contrast to

natural avalanches, skiers can directly trigger brittle fractures in buried weak layers (Föhn,

1987a; Schweizer and others, 1995).

44

2.8.2 Shear Frame

Sommerfeld (1980) and Föhn (1987a) determined correction factors for shear

frames with areas in the practical range of 0.01 to 0.05 m2. These correction factors allow

stability indices and associated critical values to be calculated and compared on the same

basis regardless of the size of the frame employed to test the weak snow layer.

However, mean strength determined with a shear frame is affected by the spacing

between fins and perhaps by the height of the fins. Little is known about the effects of

these design factors. Hence, present shear frame design is arbitrary rather than optimal.

2.8.3 Shear Frame Stability Indices

The ability of shear frame stability indices (based on the ratio of shear strength to

shear stress) to discriminate between stable and unstable slopes is evidence that slab failure

frequently begins with shear failure within the tested weak layer. Since the shear frame

stability indices rely on brittle failure of small specimens, whereas slab failure for natural

avalanches begins with a ductile failure of a very large specimen, critical levels of such

indices for natural avalanches must be determined empirically.

Slope specific stability indices such as S and S' can discriminate between stability

and instability for approximately 75% of avalanche slopes (Föhn, 1987a). Since most

accident avalanches in Canada involve failure planes that consist of persistent grain

structures, the critical levels of the stability indices for persistent snowpack weak layers

should be determined.

Extensive or frequent testing of avalanche start zones is impractical for most

backcountry forecasting operations, so extrapolated (mesoscale) stability indices need to

be calibrated for deeper layers involving persistent snowpack weaknesses.

45

2.8.4 Rutschblock

Föhn (1987b) showed that slab avalanching becomes progressively less likely as

scores from rutschblock tests on the avalanche slopes increase, and he proposed a

practical interpretation of rutschblock scores. A correlation between rutschblock scores on

safe study slopes and avalanche activity in surrounding terrain has not been established.

Jamieson and Johnston (1993b) determined the precision of 1 or 2 rutschblock scores for a

uniform slope, and showed that an increase in slope inclination of 10o tends to decrease

the rutschblock score by 1 step. This allows rutschblock scores to be estimated for steeper

avalanche starting zones from tests on less steep and safer sites nearby

46

3 METHODS

3.1 Study Areas and Co-operating Organizations

Persistent weak layers do not occur every winter in every mountain region. To

ensure sufficient data were obtained and to ensure that the results would be widely

applicable, eight locations in four different mountain ranges (Table 3.1, Figure 3.1) were

selected in co-operation with participating private or public sector organizations.

During the winters of 1992-93 to 1994-95, seasonal research technicians were based

at Bobby Burns Lodge (operated by Canadian Mountain Holidays) and at Mike Wiegele

Helicopter Skiing in Blue River, BC.

The five public sector participants were Banff National Park (BNP), Jasper National

Park (JNP), Glacier

National Park (GNP),

Yoho National Park (YNP)

and BC Ministry of

Transportation and

Highways (MoTH) at

Kootenay Pass. Wardens

from the Banff, Jasper and

Yoho National Parks, staff

from Glacier National

Park’s Snow Avalanche

Warning Section (SRAWS)

and avalanche technicians

from Kootenay Pass made

measurements

approximately once per

week during the winters of

1992-93 and 1993-94.

47

Figure 3.1 Location of study sites and mountain ranges.

The majority of the data were collected in the Cariboo and Monashee Mountains

near Blue River, BC, the Purcell Mountains near Bobby Burns Lodge, and in the Selkirk

Mountains near Rogers Pass and Kootenay Pass. These 4 ranges are part of the Columbia

Mountains where the snowpack at tree line usually exceeds 2 m in thickness throughout

the winter. Under these conditions, the average temperature gradient throughout the

snowpack is sufficiently low that thick weak layers of depth hoar are rare. Persistent weak

layers usually consist of surface hoar or faceted crystals. While buried layers of surface

hoar are usually less than 15 mm in thickness, layers of faceted crystals sometimes exceed

0.5 m in thickness. As will be shown in Section 6.4, the weak layers that cause many slab

48

Table 3.1 Study Sites and Locations

Location Main Study Sites Co-operatingOrganization

Columbia MountainsCariboo Mountains nearBlue River, BC

Mt. St. Anne Mike WiegeleHelicopter Skiing

Monashee Mountains nearBlue River, BC.

Sams Mike WiegeleHelicopter Skiing

Bobby Burns Mountains inthe Purcell Range nearParson, BC

Pygmy, Rocky, Elk,Vermont

Canadian MountainHolidays

Rogers Pass in SelkirkMountains, BC

Roundhill on Mt. Fidelity Glacier National Park,Parks Canada

Kootenay Pass in SelkirkMountains, BC

East Peak B.C. Ministry ofTransportation andHighways

Rocky MountainsBanff National Park,Alberta

Bow Summit on IcefieldsParkway

Banff National Park,Parks Canada

Jasper National Park,Alberta

Parker’s Ridge on IcefieldsParkway

Jasper National Park,Parks Canada

Yoho National Park, BC Schaffer Bowl, LakeO’Hara and Wapta Lake

Yoho National Park,Parks Canada

avalanches in the Columbia Mountains are less than 25 mm in thickness. Both natural and

skier-triggered avalanches from the Columbia Mountains are used in this study.

The balance of the data is from the Rocky Mountains where the snowpack at tree

line usually averages 1-1.5 m during the winter. Thick weak layers of faceted crystals and

depth hoar are common in the lower half of the snowpack during most winters. Almost all

avalanche data from the Rocky Mountains used in this study are for natural avalanches.

3.2 Sites for Snowpack Tests

In this study and in most forecasting operations, snowpack observations and tests

are made at avalanche start zones and at study sites. Level study sites are referred to as

study plots. Study sites are generally not threatened by avalanches except, perhaps, under

unusual conditions.

3.2.1 Study Plots and Slopes

Study sites are chosen in consultation with the co-operating organization

(Table 3.1) to ensure that general snowpack conditions at the site are representative of the

snowpack conditions common in surrounding avalanche terrain. Study sites are also

chosen for their uniform snowpack, a consideration that precludes the use of sites subject

to substantial amounts of

drifting snow. This criterion

results in a necessary difference

between study sites which are

generally sheltered from the

wind, and start zones, many of

which are exposed to drifting.

A typical study plot is shown in

Figure 3.2.

Study sites are visited

routinely, usually once or more

per week, and the snowpack tests

49

Figure 3.2 Mt. St. Anne Study Plot at 1900 m in theCariboo Mountains.

are related to avalanche activity in surrounding terrain, including avalanche activity that

may occur days after the visit to the study site.

3.2.2 Site Selection in Avalanche Start Zones

Unlike study sites

where snowpack tests are

related to avalanche

activity, often many km

away, tests in avalanche

start zones are only

related to avalanche

activity in the tested start

zone. In avalanche start

zones, snowpack tests are

made at a site judged

typical of the start zone.

This site selection requires

experience since the snowpack is usually more variable in avalanche start zones than in

study sites. Also, tests are usually completed within an area of 2-4 m2 whereas start zones

range in area from 100 to 20,000 m2. Site selection often includes probing to establish

uniformity of depth and of major layers, and sometimes preliminary profiles of the

snowpack to establish the extent of the weak layer.

Snowpack tests are made after the slope has released naturally or been ski-tested. If

a slab is successfully skier-triggered, then the area remaining for testing the undisturbed

snowpack is reduced. In some instances, no area that is typical of the undisturbed

snowpack remains after the avalanche. Field staff are shown approaching a small and large

avalanche in Figures 3.3 and 3.4, respectively.

50

Figure 3.3 Field staff approaching a small slab avalanche inthe Purcell Mountains.

3.3 Equipment

Equipment used for shear frame tests is shown in Figure 3.5.

The outer frame and intermediate fins of the shear frames consist of stainless steel or

aluminium sheet metal. Most shear frames including the standard frames used for this

study are a few mm wider at the front than at the back to reduce friction between the sides

of the frame and adjacent snow. Following experimentation with different materials and

construction techniques during 1990-92, the shear frames used from 1993-95 consisted of

a stainless steel outer frame with 0.6 mm cross fins affixed with silver solder (Figure 3.5) .

The lower edges of the frames were sharpened to reduce the force required to push the

frame into the snow above the weak layer (superstratum) and consequently to reduce the

possibility of disturbance of the weak layer.

After the frame is placed in the snow above the weak layer, a blade (Figure 3.5) is

passed around the frame to ensure that the frame is not adhering to the surrounding snow.

51

Figure 3.4 Field staff approach a 1.6 m crown fracture in the CaribooMountains for profiles and stability tests.

Force gauges with capacities of 25, 50, 100 and 250 N are used depending on the

areas of the frame and the expected strength of the weak layer to be tested. Between 10%

and 100% of their capacity, the gauges are rated accurate to within 1% of the capacity.

Setting a switch on the gauge causes it to record the maximum force.

Thermometers were used to measure the temperature of the weak layer. For most

profiles at study sites, the snow temperature was measured every 0.1 m from the surface

to below the weak layer being tested. Digital thermometers with a thermister in the tip of

the metal shaft (Figure 3.5) were used during the winters of 1993-94 and 1994-95.

The density gauge and small digital scale, or sampling tube and force gauge, were

used to measure the slab weight per unit in two different ways as described in the next

section.

For most rutschblock tests, the column of snow was isolated from the surrounding

snowpack on both sides and at the upper wall with a specialized saw (Figure 3.6). The

52

Figure 3.5 Equipment used for shear frame tests and measurement ofslab weight per unit area.

saws were all constructed of 3.2 mm aluminium and the teeth were offset to cut a 10 mm

gap in hard snow. Most of the saws were 1.3 m long and jointed so they could be

disassembled for transport by skiers with backpacks.

3.4 Measurement of Slab Weight per Unit Area

Slab weight per unit area was measured for every set of shear frame tests by one of

two methods. In the “core sample” method, the sampling tube was pushed vertically down

through the snow from the surface to the weak layer. A small plate was slid over the

bottom of the tube during extraction from the snowpack. The snow in the tube was then

deposited into a light plastic bag. If the slab height exceeded the length of the tube,

subsequent samples were taken directly below the previous sample until the weak layer

was reached. To obtain an average weight, the procedure was usually repeated several

times before the bag was weighed by suspending it from a vertically held force gauge. For

thin slabs, additional samples were required to ensure the weight of the core samples

exceeded 10% of the force gauge’s capacity. Dividing the weight by the 0.0028 m2

cross-sectional area of the tube and by the number of cores yields the slab weight per unit

area, denoted by ρgh in this dissertation. Avalanche workers usually refer to the slab

weight per unit area as the load on the weak layer.

Figure 3.6 Rutschblock saws.

53

The second method of measuring load requires that the layers be identified. A

sample of each layer was taken with a 93 mm long cylindrical sampler with a diameter of

37 mm and a volume of 1 x 105 mm3. The sampler was inserted vertically into layers

thicker than 0.1 m, and horizontally into thinner layers. These samples were either

weighed using the gravity balance or by placing the snow in a plastic bowl on the small

digital scale (Figure 3.5) which had been tared with the empty bowl. The slab weight per

unit area is simply the sum over the layers of the product of density, layer thickness and

acceleration due to gravity.

The advantages of the layer-by-layer density method are:

field workers can mentally calculate the density from the weight (by shifting the

decimal point) and repeat the sample if the density seems questionable, and

the density of the individual layers can be correlated with other properties such as

resistance, as was done for hand hardness (Appendix A).

The core sample method is faster and potentially more accurate since the volume of

snow that is weighed is larger. However, two accuracy problems were identified with the

core sample method. Sometimes on slopes operators do not hold the tube vertically before

pushing it down through the slab. This can be mitigated if a second person, standing back

several metres, carefully watches and corrects the orientation of the tube. The second

problem occurs when crusts overlie soft snow layers within the slab. Under these

conditions, the descending tube will often break the crust into pieces larger than the

cross-section of the tube, and the softer snow below will be pushed ahead of and away

from the descending tube, resulting in under-sampling.

When time permitted and crusts were not a problem, both methods were used and

the results averaged.

3.5 Shear Frame Tests

The main advantages of the shear frame test method are:

it is conceptually simple,

the strength of very thin layers can be determined,

54

the frame and force gauge are small enough and light enough to be carried in a

backpack in the field, and

the frame and force gauge are relatively inexpensive (< $800).

Its disadvantages include:

placing the frame on thin, often delicate weak layers requires considerable manual

skill,

weak layers underlying a crust often cannot be tested since pushing the frame

through the crust may fracture the weak layer,

the results depend on the design of the frame (Perla and Beck, 1983), and

load is applied manually and the results depend on the loading rate (Perla and Beck,

1983; Föhn, 1987a).

Studies of these last three limitations are discussed under Size Effects

(Section 4.10), Frame Design (Section 4.12) and Loading Rate (Section 4.5). Variability

between operators is discussed in Section 4.9.

3.5.1 Technique

Before the shear frame test is performed, the weak layer is identified with a profile

of snow layers, a tilt board test, a shovel test, or a rutschblock test, all of which are

described in Observation Guidelines and Recording Standards for Weather, Snowpack

and Avalanches (CAA, 1995). Overlying snow is removed, leaving approximately

40-45 mm of undisturbed snow above the weak layer (Figure 3.7).

The shear frame with sharpened lower edges is then gently inserted into the

undisturbed snow so that the bottom of the frame is—preferably—within 2-5 mm, of the

weak layer (Perla and Beck, 1983). In practice, the strengths of the weak layer and the

snow above the weak layer (the superstratum) influence the distance between the weak

layer and the bottom of the frame. If the superstratum is not much harder than the weak

layer then the shear frame must often be placed very close to, or into, the weak layer to

avoid a fracture in the superstratum rather than in the weak layer when the frame is pulled.

Alternatively, if the superstratum is very hard, then the weak layer may pre-fracture, that

55

is, fracture during frame placement.

Under such conditions, frames must be

placed 5-10 mm above the weak layer

and occasionally higher. The effect of

frame placement more or less than the

recommended 2-5 mm is discussed in

Sections 4.8 and 5.4. After the frame is

placed, a thin blade is passed around the

sides of the frame to ensure that

surrounding snow is not in contact with,

and possibly bonding to, the frame. This

cut must extend to the weak layer to

ensure that a known area is tested. The

force gauge is attached to the cord

linking the two sides of the frame and is

pulled smoothly and quickly (< 1 s)

usually resulting in a planar failure in the

weak layer just below the base of the

frame. Tests in which half or more of the fracture surface deviated beyond the active weak

layer were rejected. Shapes of fracture surfaces and treatment of data with non-planar

fracture surfaces are discussed in Section 4.4. Shear strength is determined by dividing the

maximum load on the force gauge by the area of the frame, usually 0.025 m2.

Average shear strengths of weak layers were based on sets of at least 7 shear frame

tests during the winters of 1990-93. During the winters of 1994 and 1995, the usual

number of tests in study sites was increased to 12 to reduce the standard error of shear

strength.

56

Figure 3.7 Shear frame test. (J. Hughesphoto)

3.6 Rutschblock Test

3.6.1 Rutschblock Technique

A pit at least as deep as

any potential failure

planes—often 1-1.5 m deep—is

excavated with a shovel. The

wall of the pit that faces

down-slope is extended by

shovelling until it is at least 2 m

across the slope (Figure 3.8).

The sides of the block can be

either cut or shovelled, the latter

method requiring more time. If the sides of the block are to be shovelled, then two 1.5 m

long parallel marks extending up the slope from the pit wall and 2 m apart are made on the

snow surface with a ski or ruler. After shovelling trenches just outside these marks, the

upper wall is cut with the tail of a ski

or a cord.

If the side walls are to be cut

with a ski, ski-pole, cord or saw, then

the marks for the side walls are 2.1 m

apart at the pit wall and 1.9 m apart at

the up-slope end of the marks

(Figure 3.9). This flaring of the block

reduces the potential for friction at the

sides of the block that could affect the

rutschblock score.

After the side walls are cut with

a ski-pole, saw or tail of a ski by a

person standing outside the

57

Figure 3.9 Rutschblock isolated on the sides andupper wall by cord cutting.

Figure 3.8 Rutschblock isolated on the sides byshovelled trenches.

dimensions of the column, then the upper wall is cut by the same method. However, if a

cord is used, then the side walls and upper wall can be cut simultaneously by extending the

cord from the pit, up one side of the column, around ski-poles or avalanche probes at the

upper corners and down the other side to the pit (Figure 3.9). Two operators in the pit,

one holding either end of the cord, alternately pull their end of the cord to “saw” both side

walls and the upper wall. An 8 m length of 4-6 mm cord with simple knots tied every

0.3 m successfully cuts a wide variety of snow layers except for melt-freeze crusts of

“knife” hardness.

3.6.2 Loading Steps and Rutschblock Scores

Rutschblock scores range from 1 to 7. Scores of 1-6 correspond to the first loading

step that produces a slope-parallel failure of the block. A score of 7 indicates that none of

the 6 loading steps caused a slope-parallel failure. The following sequence of loading

steps, except for the “soft slab” variation of step 6, is similar to the steps described by

Föhn (1987a):

Step 1: An undisturbed column of snow is isolated by shovelling or cutting as

described above.

Step 2: The skier approaches the block from above and gently steps down onto the

upper part of the block (within 0.35 m of the upper wall).

Step 3: Without lifting the heels, the skier drops from a straight leg to a bent knee

position, pushing downwards and compacting surface layers.

Step 4: The skier jumps upwards, clear of the snow surface, and lands on the

compacted spot.

Step 5: The skier jumps again and lands on the same compacted spot.

Step 6: For hard or deep slabs, the skier removes the skis and jumps on the same spot.

For softer slabs where jumping without skis might penetrate through the slab,

the skis are kept on, the skier steps down another 0.35 m—almost to

mid-block—and pushes downwards once then jumps at least three times

58

3.6.3 Failure Mode

For rutschblock scores of

1 or 2 and sometimes for

scores of 3 or higher, the entire

block displaces as shown in

Figure 3.10. However, when

the loading steps of 4, 5 or 6

are applied to softer slabs, the

fracture often extends from the

operator’s skis down to the

weak layer and along the weak

layer to the pit, leaving a part

of the block undisplaced. In

such cases, the area of the

shear failure is less than 2 m2.

However, this reduction in area

is minimized by loading the

rutschblock near the upper wall

as shown in Figures 3.8, 3.9

and 3.10.

3.7 Comparison of Rutschblock and Shear Frame Tests

The rutschblock method can test a 2 m2 area of snowpack in 10-20 minutes whereas

a set of 7-12 shear frame tests in a 2 m2 pit typically requires 30-45 minutes, including slab

weight measurements.

A rutschblock test can be interpreted immediately whereas computing a stability

index from shear frame tests requires a hand-held calculator or written calculations.

Further, a displaced rutschblock is often a convincing indication of instability whereas a

shear frame stability index is “just a number”.

59

Figure 3.10 Rutschblock showing displaced block(M. Shubin photo)

Although site selection for any stability test requires experience, the rutschblock test

requires less practice and less specialized equipment (the large saw is optional) than the

shear frame test.

The shear frame test can only be used after the weak layer has been identified,

whereas the rutschblock test identifies weak layers and rates their stability. However, the

“ski-penetration problem” may cause the rutschblock test to overlook weak layers or

incorrectly rate their stability. This problem occurs when the operator's skis penetrate very

close to, or through, weak layers during the loading steps (Jamieson and Johnston,

1993b).

After a shallower slab has displaced, allowing the operator's skis to penetrate the

slab more deeply and increasing the skier-induced stress, the rutschblock test may not

yield a valid score for deeper weak layers. This can be a substantial limitation if the

stability of deeper weak layers is important, or as was often the case in the present study,

the primary objective of the test.

The rutschblock test is not a reliable test for weak layers deeper than 1 m whereas

the shear frame test can be applied at any depth.

Jamieson and Johnston (1993b) found a significant effect of slope inclination on

rutschblock score only for 10 of 24 slopes and Schweizer and others (1995) doubt that a

rutschblock score can be adjusted for slope inclination due to the non-linear loading steps.

This is in contrast to shear frame stability indices which can calculated for a wide range of

slope inclinations. Not only can the slope inclination be changed, but the slab density and

slab thickness can be readily changed to estimate the stability where conditions are

different.

Shear frame stability indices for skiers are calculated for a static skier and may not

represent the dynamic shear stress due to skiing or, in the worst case, the impact of a

falling skier. On the other hand, the same set of shear frame and slab weight measurements

can be used to calculate stability indices for triggering due to slab weight (natural trigger),

a skier, over-snow vehicle, explosive, etc.

60

The stiffness of superstratum influences slab failure by imposing stress and strain

concentrations on the weak layer. The rutschblock test includes this real stress and strain

concentration whereas the shear frame test obscures this effect by imposing a stress

concentration due to the fins and dimensions of the frame on the weak layer. Also, pushing

the frame into a hard superstratum sometimes fractures the weak layer, making valid shear

frame tests impossible.

Selecting the best test method based on the comparisons summarized above depends

on the objective. The present research study uses both rutschblock and shear frame tests in

avalanche start zones and in study sites. However, for avalanche forecasting programs,

some general preferences can be summarized. Both the shear frame test and rutschblock

test are well suited to study sites. However, the shear frame stability indices have

advantages for extrapolation to surrounding terrain since they can be calculated for

different slope inclinations, slab densities and slab thicknesses as well as for different

triggers (natural, explosive, etc.). The ability to calculate shear frame stability indices for

different triggers is less important for forecasting programs concerned primarily with

skier-triggered avalanches. While stability tests at study sites are applied to surrounding

terrain, tests in avalanche start zones are more often applied locally. For many forecasting

programs, the speed and simplicity of the rutschblock test are substantial advantages in

start zones. Based on the literature summarized in Table 2.2, operational use of the shear

frame—which dates back to 1963—has been restricted to study sites.

3.8 Avalanche Activity

For the present study, avalanche occurrences were compiled by type of release (slab

or loose), size, type of trigger (natural, cornice or skier-released, etc.), moisture content

(dry, moist or wet), aspect, elevation and location (CAA, 1995) using mainly information

obtained from ski guides operating in the forecast area. On a given day the portion of the

total operating area observed for avalanche occurrences varied from 0 to 40% depending

on visibility conditions, number of guides skiing (typically 5-12) and their operating

locations. The research team also compiled occurrence data for slopes visible from near

61

the study sites, in particular on days during bad weather when helicopter skiing operations

were grounded.

Some avalanche occurrence data are unavoidably influenced by weather and

operational factors. Typically, this happened when, for one or more days after an

occurrence, visibility was limited or there was no helicopter skiing near the location of the

avalanche. Some crown fractures and/or deposits from natural avalanches were estimated

to be 1 or more days old when they were first observed. Consequently, for most of these

avalanches, the date of occurrence was estimated. In a determination of critical values of

stability indices, Jamieson and Johnston (1993a) found that excluding avalanches with

estimated dates had little effect on the critical values. In Chapters 6 and 7, the stability

parameters are related to natural avalanche activity including avalanches with estimated

dates. Fortunately, for skier-triggered avalanches the dates of occurrence are known.

For the purpose of the present study, natural avalanches are defined as those that

release without an external trigger such a skier, explosive or falling chunk of cornice.

Cornice-triggered avalanches are not considered natural avalanches in the present study

since many falling chunks of cornice are powerful triggers that can release relatively stable

slabs. Distinguishing cornice-triggered avalanches from natural avalanches is consistent

with the NRCC/CAA (1989) definition of natural avalanches and with Roch’s (1981)

definition of an intrinsic trigger, but inconsistent with the CAA (1995) definition. Slab

avalanches triggered by explosives and helicopters were recorded but there were too few

of these avalanches to use in the analysis.

In Canada, avalanches are classified by size based on destructive potential

(Table 3.2; CAA, 1995). A class 1 avalanche is “relatively harmless to people”, whereas a

class 2 avalanche can “injure, bury or kill a person”. A class 3 avalanche can “bury and

destroy a car, destroy a small building, damage a truck or break a few trees”. The

destructive potential for larger avalanches is given in Table 3.2. Half sizes, such as 1.5, are

used for avalanches that appear to fall between two size classes.

Most reported avalanches were within 15 km of the study sites, but some were more

than 30 km away.

62

Table 3.2 Avalanche Size Classification (CAA, 1995)

Size Class Destructive Potential

1 Relatively harmless to people.

2 Could injure, bury or kill a person.

3 Could bury and destroy a car, destroy a small building, damage atruck or break a few trees.

4 Could destroy a railway car, large truck, several buildings, or aforest area up to 4 hectares.

5 Could destroy a village or 40 hectares of forest.

63

64

4 FIELD STUDIES OF THE SHEAR FRAME TEST

4.1 Introduction

Shear strength, as measured with a shear frame, depends on the material properties

of the snow layer being tested. However, the results of shear frame tests may also be

affected by or related to factors such as:

manual loading rate,

different operators,

test sequence since the operator may refine the manual loading rate or the frame

placement while repeatedly testing the same layer,

delays between placing the frame and applying the load,

shape of the fracture surface,

normal load on the weak layer,

distance between the bottom of the frame and the weak layer,

cross-sectional area of frame, and

design of frame including height of frame and number of load-carrying

cross-members.

The effects of these factors are assessed in this chapter after examining the

statistical distribution and variability of shear strength measurements.

4.2 Statistical Distribution

Strengths determined from shear frame tests can be considered continuous, and

certainly they have the properties of interval data. Analyses of such data are facilitated if

the data are normally distributed.

To assess the normality of shear frame results, 28 sets of 30 or more tests with

standard frames from the winters of 1991-95 are summarized in Table 4.1. The

Shapiro-Wilk test is presently the preferred test for normality (Shapiro and others, 1968;

Statsoft, 1994, p. 1412). For four of the 28 sets of shear frame tests tabulated in

Table 4.1, the hypothesis of normality is rejected at the 1% level (p < 0.01). An additional

four sets could be rejected at the 5% level (p < 0.05). The values of p for the 28 large sets

65

66

Table 4.1 Normality of Large Sets of Shear Frame Tests

Date Microstructure(of most common

grains in weak layer)

MeanStrength

(kPa)

Coef.of

Var.V

No. ofTests

n

Shapiro-Wilk

Test1

W p

91-12-19 decomposed & frag. 0.301 0.167 30 0.891 0.002

91-12-20 precip. particles 0.345 0.100 36 0.984 0.913

91-12-21 decomposed & frag. 0.525 0.084 32 0.913 0.015

92-02-11 decomposed & frag. 1.096 0.108 38 0.824 1x10-5

92-02-14 surface hoar 0.386 0.083 30 0.931 0.059

92-02-17 surface hoar 0.645 0.129 30 0.958 0.320

92-04-10 facets 0.614 0.195 32 0.972 0.615

93-02-06 surface hoar 2.129 0.199 32 0.987 0.963

93-02-13 surface hoar 3.181 0.144 32 0.970 0.568

93-02-24 surface hoar 0.476 0.141 32 0.948 0.150

93-03-03 surface hoar 0.615 0.147 30 0.962 0.394

93-03-16 facets 2.157 0.172 31 0.940 0.098

93-04-01 facets 2.208 0.179 30 0.979 0.824

94-03-30 graupel 4.013 0.123 30 0.952 0.222

94-12-04 graupel 0.931 0.135 30 0.973 0.673

94-12-15 decomposed & frag. 0.219 0.135 30 0.908 0.014

94-12-29 graupel 1.432 0.197 30 0.891 0.005

95-01-04 graupel 2.124 0.160 30 0.950 0.201

95-01-29 surface hoar 1.294 0.199 30 0.921 0.033

95-02-09 surface hoar 3.71 0.090 33 0.888 0.002

95-02-22 surface hoar 1.874 0.118 30 0.947 0.160

95-02-28 surface hoar 3.035 0.083 36 0.968 0.457

95-03-07 surface hoar 4.185 0.075 30 0.970 0.594

95-03-24 graupel 2.764 0.150 30 0.923 0.038

95-03-28 graupel 3.597 0.10 30 0.977 0.770

95-03-29 surface hoar 5.921 0.09 30 0.963 0.408

95-01-23 surface hoar 4.139 0.097 30 0.964 0.444

95-01-29 surface hoar 1.204 0.109 33 0.962 0.3521 Rows for which p ≤ 0.05 are marked in bold.

of shear frame tests are plotted against mean shear strength in Figure 4.1 using

international symbols (Colbeck and others, 1990) to distinguish the microstructure. No

systematic effect of mean shear strength on p is apparent in Figure 4.1. However, p < 0.05

for all four sets of tests on decomposed and fragmented precipitation particles. The

frequency histograms and associated normal distributions for the eight sets for which

p < 0.05 are presented in Figure 4.2. All exhibit central tendency; however, four are

skewed right and two are skewed left. Since the hypothesis of normality cannot be

rejected for 20 (p < 0.05) to 24 (p < 0.01) of the 28 data sets, the shear frame data are

assumed to be normally distributed. However, further studies of decomposed and

fragmented precipitation particles would be worthwhile.

Figure 4.1 Shapiro-Wilk test for normality for 28 sets of shear frame data.

67

68

Figure 4.2 Frequency distributions for 8 sets of shear frame tests for whichp < 0.05 from Shapiro-Wilk test for normality.

4.3 Variability and Number of Tests for Required Precision

The spatial variability of shear strength of a particular weak layer throughout an

area such as an avalanche start zone is certainly relevant to stability evaluation, but digging

numerous pits in every start zone or study plot of concern requires too much time for any

avalanche forecasting operation. Safety concerns also limit access to avalanche start

zones. Hence, the variability reported for the present studies and for previous studies

(except for Sommerfeld and King, 1979) is for repeated shear frame tests of a particular

weak layer in a single pit, usually within 30-60 minutes. Shear frame tests are usually

made in one or two rows across the slope and the area tested by 7 to 12 shear frame tests

is usually less than 0.5 m by 2 m.

The coefficient of variation is the preferred measure of variability for snow

strength since it is less dependent on mean strength than the standard deviation which

increases with mean strength (Keeler and Weeks, 1968; Jamieson, 1989). Previously

reported coefficients of variation for 0.01, 0.025 and 0.05 m2 shear frame tests are

typically 0.25 but range from 0.18 to 0.62 (Perla, 1977; Sommerfeld and King, 1979;

Föhn, 1987a; Schaerer, 1991). The highest values of 0.54, 0.52 and 0.62 were reported by

Sommerfeld and King (1979) who tested the failure plane at up to three sites along crown

fractures.

During the winters of 1989-90 to 1994-95, 809 sets of six or more shear frame

tests (with areas of 0.01, 0.025 or 0.05 m2) were made in study sites chosen partly for

uniformity of snowpack (Section 3.2.1), and in avalanche start zones where greater

variability is expected. For these 809 sets, the coefficients of variation ranged from 0.03 to

0.66 with a mean of 0.152 (Figure 4.3). At level study plots (inclination <3°) chosen for a

uniform snowpack, the coefficients of variation from 342 sets of six or more tests ranged

from 0.03 to 0.46 with a mean of 0.144. In avalanche start zones with inclinations of at

least 35°, the coefficients of variation from 114 sets of six or more tests ranged from 0.04

to 0.54 with a mean of 0.178. The remaining 353 sets are from sites, mainly study slopes,

with slope inclinations between 3° and 34°.

69

The present study’s coefficients of variation are generally lower than previous

studies. This may be due to:

reduced disturbance of the weak layer due to the use of relatively shear frames

constructed of relatively thin sheet metal with sharpened lower edges,

consistently fast loading resulting in fractures within 1 s,

the generally consistent snowpack of the Columbia Mountains where most of the

tests were done, or

the field practice of rejecting tests that do not fail on the intended layer, or show

definite evidence of disturbance such as pine needles or an animal track in the

fracture surface. (Tests were not rejected due to surprisingly low or high pull

forces.)

The number of tests, n , required to obtain precision, P , can be estimated from the

coefficient of variation, V, by solving

n = (tp;n-1 V/P)2 (4.1)

Table 4.2 shows the number of tests for V = 0.15 which is typical of study plots in

the present study, V = 0.18 which is typical of start zones in the present study and

70

Figure 4.3 Frequency distribution for coefficients of variation of shear strength.

V = 0.25 which is typical of previous studies. For any particular level of precision, the

number of tests increases with the coefficient of variation. The reduced variability from the

present study results in fewer tests being necessary to achieve a required precision. This an

important point for avalanche safety operations because a greater number of tests would

require a larger pit and more time, and could be operationally impractical.

4.4 Fracture Surface

During the winters of 1991-95, 703 sets of shear frames tests were made with

standard 0.025 m2 frames for a total of 8468 tests. The shape of the fracture surfaces were

classified with the standard descriptors (Table 4.3) for 6951 tests, and 5757 of these

(83%) exhibited smooth planar fracture surfaces. Shapes that occurred less than 10 times

are excluded from Table 4.3 and this analysis.

To compare the strength measurements associated with planar fracture surfaces

with measurements associated with other shapes of fracture surfaces, each strength, Σi,

was converted to standard normal

(4.2)ui = (Σ i − Σ)/sΣ

where and sS are the mean and standard deviation for the particular set of shear frameΣ

71

Table 4.2 Number of Shear Frame Tests for Required Precision

RequiredPrecision

P

SignificanceLevel

p

No. of Shear Frame Tests

Coef. of Var.0.15

Coef. of Var.0.18

Coef. of Var.0.25

0.05 0.10 26 37 70

0.05 0.05 36 52 98

0.05 0.01 64 90 170

0.10 0.10 8 11 19

0.10 0.05 11 15 26

0.10 0.01 19 25 45

0.15 0.10 5 6 9

0.15 0.05 6 8 13

0.15 0.01 10 13 22

tests. Combining the normalized values from the 703 sets gives an aggregate set of 8468

values with mean 0 and standard deviation 0.95. This aggregate set is then partitioned by

fracture descriptor so the set of values with a particular fracture descriptor can be

compared with the set of 5757 values with planar fractures using a two-tailed t-test for

unequal sample sizes and unequal variances. Representing the sets of planar and

non-planar fracture surfaces with the subscripts 1 and 2 respectively, the calculated value

for t (Mattson, 1981, p. 430) is

(4.3)t = (u1 − u2)/(s12/n1 + s2

2/n2)1/2

72

Table 4.3 Assessment of Common Shapes of Fracture Surfaces

Sample t-test

Descriptor Description of FractureSurface

n Mean

(kPa)

St.Dev.(kPa)

t p

C smooth, planar 5757 -0.02 0.95 - -

SBD divot under rear compartment,divot < 5 mm deep

200 0.02 0.93 -0.66 0.51

MBD divot under rear compartment,divot 5-10 mm deep

154 0.09 0.90 -1.46 0.15

BBD divot under rear compartment,divot > 10 mm deep

131 0.41 0.93 -5.19 7E-07

W 1 wave per compartment, waves 5-10 mm deep

243 0.01 0.94 -0.54 0.59

SW 1 wave per compartment, waves < 5 mm deep

77 0.03 0.85 -0.54 0.59

SIR irregularities < 5 mm deep 44 0.13 0.9 -1.09 0.28

IRR irregularities 5-10 mm deep 114 0.10 0.96 -1.28 0.20

LC fractured deeper at right orleft side

31 0.27 0.87 -1.84 0.07

SH small hump, height < 5 mm 22 -0.25 1.04 1.04 0.31

STP stepped between 2 fractureplanes

21 0.31 0.94 -1.60 0.12

BC back divot extends beyondrear compartment

14 -0.11 1.30 0.26 0.80

where the number of degrees of freedom is

df = (se12 + se2

2)2 / (se14/n1+se2

4/n2) (4.4)

and the standard error is

(4.5)se = s/ n

For each comparison, calculated values of t and the total probability associated

with both tails, p, are shown in Table 4.3.

The only shape of fracture descriptor that has strength measurements significantly

different from planar fractures are those with a back divot under the rear compartment

that is more than 10 mm deep. Such fractures only occur when the bed surface is not

appreciably stronger than the weak layer being tested, typically when both the weak layer

and the bed surface consist of facets or depth hoar. This condition is most common in the

Rocky Mountains where entire sets of shear frame tests may result in fractures with deep

back divots. For future studies in which minimal variability is important, shear frame tests

with such fracture surfaces should be rejected. However, this rejection may not be

practical in the Rocky Mountains where deep back divots are common and imperfect data

may be better than none for avalanche forecasting.

73

Table 4.4 Mean Shear Strength for Various Loading Times

Date Microstructure Loading Time to Failure (s)

0-0.5 0.6-1.0 1.5-3.0 3.5-6.0 6.5-9.0 9.5-15 15.5-30

90-01-28 decomposed/frag. - 0.37 0.34 0.44 - - -

90-01-20 rounded grains 0.35 0.38 0.39 0.39 0.32 - 0.41

90-01-28 decomposed/frag. - 0.4 0.43 0.34 - - -

90-01-20 rounded grains 0.42 0.44 0.5 0.53 0.58 0.51 -

90-01-21 decomposed/frag. 0.51 0.52 0.51 - 0.6 - -

90-01-28 rounded grains - 1.11 1.2 1.31 1.48 - -

90-01-29 decomposed/frag. - 1.23 1.45 1.51 1.38 1.8 -

90-01-22 rounded grains - - 1.98 1.91 2.07 - -

90-01-21 rounded grains 1.75 2 2.09 2.33 - 2.54 -

90-01-20 rounded grains 2.32 2.9 2.87 3.23 2.83 3.1 4.7

4.5 Loading Rate

In a laboratory study

of depth hoar under various

constant displacement rates,

Fukuzawa and Narita (1993)

found a ductile-brittle

transition between 8 x 10-5 s-1

and 2 x 10-4 s-1. In

ductile-brittle transition, as

loading times to failure

decreased from 3 to 1.5 s,

strengths were reduced by

approximately 35%.

However, in the brittle range as

loading times decreased from 1.5 s

to 0.2 s, strengths were only

reduced a further 12%.

For in situ shear frame tests of snowpack layers, the load is applied manually with

a force gauge that records the maximum force. Shear strength measured with a shear

frame depends on the rate at which the manual load is applied. Perla and Beck (1983)

found the strength was reduced by 25% when loading times were reduced from

approximately 30 s to approximately 3 s. To minimize rate effects, recent field studies

(Föhn, 1987a; Jamieson and Johnston, 1993a) chose to load shear specimens to failure

within 1 s.

The effect of loading rate on shear strength was studied by attempting to apply the

load at various constant rates and measuring the loading time with a stop-watch. Loading

times are partitioned into 7 intervals and for each interval in which there were at least

three results, the mean strength is reported in Table 4.4 and plotted in Figure 4.4. For

strengths less than 1 kPa, shown in the first five rows of Table 4.4, there is no apparent

74

Figure 4.4 Effect of loading time on shear strengthfor 10 experiments with various manual loadingrates.

effect of loading time on strength. However for four of the five series with strengths

greater than 1 kPa (shown in the last 5 rows of Table 4.4), there is an increase in strength

with a increase in loading times (decrease in loading rate).

Since persistent weak layers typically have a shear strength of less than 2 kPa and

critically weak persistent layers typically have a strength less than 1 kPa, shear frames

were loaded to failure in less than 1 s for all shear frame results presented in subsequent

chapters and in the other field studies of the shear frame test outlined in this chapter. This

minimizes the effect of loading rate on strength and is consistent with the laboratory study

of Fukuzawa and Narita (1993).

4.6 Test Sequence Variability

Schaerer (personal communication) has suggested that the first shear frame test in

a set of tests on a particular weak layer be rejected since the operator requires at least 1

test to learn the optimal loading rate and frame placement with respect to the weak layer.

To determine if variability is greater for initial tests than for subsequent tests, 703

sets of 2 or more tests with the standard 0.025 m2 frame are used. The strength from each

test is normalized using the mean and standard deviation from each set (Eq. 4.2). The

combined set of normalized strengths is then partitioned by sequence number. The means

and standard deviations of the set of first tests, set of second tests, etc. and including the

set of third to seventh tests are shown in Table 4.5. The standard deviations of these sets

are plotted in Figure 4.5.

Increased variability on the first and second tests in a sequence is apparent in

Figure 4.5. The significance of the apparent increase in variability is assessed by

comparing the variance of the set of first tests with the variance of the set of third to

seventh tests with an F test.

F = (s1/s3-7)2 (4.6)

This value of F is 1.45 which is significant at the 10-4 level. Similarly, when the variance of

the set of second tests is compared with the variance of the set of third to seventh tests,

F = 1.14 which is marginally significant (p = 0.03). Hence for future studies in which

75

minimal variability is important, the first and perhaps the second test in a set should be

rejected.

4.7 Effect of Delay

Shear frames are usually pulled within 5-10 s of the frame being placed in the snow

above the weak layer. Occasionally the frame remains placed for up to 60 seconds while

the force gauge is being zeroed, cleaned of snow, ice or moisture. To assess the effect of

Table 4.5 Effect of Test Sequence on Variability

SequenceNo.

No. of Testsin Set

Mean St. Dev

1 703 -0.004 1.111

2 703 -0.008 0.985

3 702 0.007 0.903

4 701 0.06 0.943

5 696 0.001 0.91

6 691 0.048 0.92

7 674 0.005 0.93

3-7 3464 0.024 0.921

Figure 4.5 Effect of sequence number on standard deviation.

76

such delays, alternating tests were performed on 30 March 1995 at the Mt. St. Anne Study

Plot. This weak layer was susceptible to changes in strength because the weak layer was

relatively warm (-3.7oC), the air was unusually warm (+7oC) and the microstructure was

non-persistent (partly decomposed precipitation particles and rounded grains) and

consequently capable of metamorphic and strength changes faster than most layers in the

present study.

Tests with the usual 5-10 s delay were alternated with tests with a 3 minute delay

until 14 pairs were obtained. The hypothesis that there was no difference in strength was

assess with a two-tailed t-test for matched pairs

(4.7)t = D n /sD

where and sD are the mean and standard deviation of the differences in strengthD

respectively.

The comparison was repeated with a 3 kg mass on top of the shear frame. In each

case, there was no significant difference in strength (p > 0.14) as shown in Table 4.6.

Apparently, delays of up to 3 minutes do not affect the strength measured with a shear

frame. This would also be true for persistent layers, the mechanical properties of which

are, by definition, slower to change than non-persistent layers.

4.8 Frame Placement

When a shear frame is loaded, stress concentrations occur in the snow at the lower

edges of the cross-members. Placing the lower edges closer to a weak layer should result

in lower strength measurements due to increased stress concentrations in the weak layer.

Recommended distances between the bottom edges and the weak layer include “< 5 mm

Table 4.6 Effect of Delay on Shear Strength

Load on Topof ShearFrame (kg)

No Delay 3 Min. Delay Difference t-test for PairedTests

Mean

(kPa)

Coef. ofVar.

Mean

(kPa)

Coef. ofVar.

Mean

(kPa)

Coef. ofVar.

No. ofPairs t p

0 2.74 0.13 2.73 0.18 -0.01 39.5 14 0.09 0.926

3 3.59 0.05 3.41 0.11 -0.18 2.04 14 1.56 0.143

77

but not through the weak layer” (Perla and Beck, 1983), “just above” (Sommerfeld, 1984;

Schaerer, 1991), “a short distance above” (NRCC/CAA, 1989), and “a few mm above”

(CAA, 1995).

To assess the effect of the distance between the lower edges of the frame and the

weak layer, four sets of 14 to 33 alternating pairs were made during the winter of 1994-95

with a standard 0.025 m2 frame. Each pair consisted of a test with the lower edges placed

in the weak layer and one with the lower edges placed above the weak layer. When the

lower edges were placed in the weak layer the strength was reduced by 12%, 13% and

20% compared to strengths obtained with the lower edges 2-5 mm above the weak layer

(Table 4.7). This strength reduction was 41% when the lower edges were placed in the

weak layer compared to strengths obtained with the lower edges 10 mm above the weak

layer. In every comparison, the strength reduction was significant (p < 10-5) based on

two-tailed t-test for matched pairs (Eq. 4.7).

To reduce stress concentrations, it is clearly advantageous to place the frame

above the weak layer. However, this is not always practical since it will sometimes result

in a failure within the snow above the weak layer (superstratum). This is most common

when the superstratum is comparable in strength to the weak layer. Resulting fracture

surfaces may be “wavy” or “irregular”. When failures occur in the snow above the weak

78

Table 4.7 Effect of Frame Placement on Shear Strength

Date /Microstructure

Above Weak Layer In WeakLayer

Difference t-test for Pairs

Dist.above(mm)

Mean

(kPa)

C. ofVar.

Mean

(kPa)

C. ofVar.

Mean

(kPa)

C. ofVar.

No. ofPairs

t p

94-12-16 rounded facets

10 0.80 0.16 0.46 0.24 -0.33 -0.54 14 6.9 1x10-5

95-01-29surface hoar

2-5 1.29 0.20 1.04 0.17 -0.26 -0.95 30 5.7 3x10-6

95-1-23surface hoar

2-5 4.14 0.11 3.65 0.13 -0.49 -1.14 30 4.8 4x10-5

95-1-29surface hoar

2-5 1.20 0.10 1.06 0.11 -0.15 -1.00 33 5.7 2x10-6

layer, usually the only feasible way to test the weak layer is to place the lower edges of the

frame in the weak layer.

Fortunately, weak layers of surface hoar commonly result in planar fractures when

the shear frame is placed above the weak layer. Occasionally, the superstratum is so hard

that it fractures prematurely when the operator pushes the frame into the weak layer. Such

“pre-fractures” can sometimes be avoided by pushing the frame only to within 10-20 mm

of the weak layer. These “high” placements may result in higher strengths due to reduced

stress concentrations in the weak layers or, conceivably, lower strengths due to bending.

However, there is no evidence of reduced strength in the one set of shear frames placed

10 mm above the weak layer (Table 4.7). The sensitivity of shear strength to the distance

that the frame is placed above the weak layer represents a limitation of the shear frame

test.

Until there are further studies of this effect, it is recommended that, whenever

possible, the shear frame should be placed 2-5 mm above the weak layer, and that

whenever snowpack conditions necessitate that the distance between the frame and the

weak layer be more or less than the nominal 2-5 mm, the distance should be recorded.

In subsequent chapters, results based on shear frames placed more or less than the

recommended 2-5 mm above the weak layer are included.

4.9 Variability Between Operators

The results of shear frame tests can vary between operators because the frame is

placed manually and the load is applied manually. Also, operators vary in their ability to

locate very thin weak layers but their skills improve with training and experience over one

or more winters.

Operator variability was studied by alternating operators while testing the same

layer on the same day. Using the difference in strength measurement between adjacent

tests with alternating operators, D, the hypothesis that is tested with a two-tailedD = 0

t-test (Eq. 4.7). The results of 20 comparisons are summarized in Table 4.8 which includes

a column for the mean loading time since many of these sets of shear frame tests involve

79

loading times greater than 1 s. The hypothesis was rejected at the 1% level (p < 0.01) for

one comparison involving 22 pairs and for two additional experiments at the 5% level

(p < 0.05). In the comparison that a significant difference was detected (p < 0.01), one

operator was tapping the frame into place with the blade used to cut around the frame

while the other was using the standard technique of pushing it into place by hand. When

80

Table 4.8 Effect of Different Operators on Shear Strength

Date Micro-structure

MeanLoadTime

(s)

Operator 1 Operator 2 Difference t-test

MeanStrength± S.D.(kPa)

MeanStrength

±S.D.(kPa)

Mean

(kPa)

Coef.of

Var.

No.of

Pairst p

90-01-15 decomp./frag. 1.4 1.10±0.10 1.19±0.23 -0.09 -1.9 7 1.38 0.218

90-01-15 graupel 1.7 3.40±0.39 3.01±0.31 0.38 1.6 9 1.87 0.099

90-01-31 rounded 1.3 1.69±0.31 1.60±0.22 0.09 3.8 8 0.74 0.485

90-03-02 decomp./frag. 1.0 1.64±0.24 1.58±0.18 0.06 4.8 17 0.86 0.403

90-03-02 decomp./frag. 1.3 4.12±0.47 3.90±0.50 0.21 1.9 19 2.30 0.034

90-03-04 decomp./frag. 0.9 2.23±0.23 2.11±0.28 0.12 2.0 22 2.29 0.032

90-03-04 decomp./frag. 1.0 2.07±0.17 2.11±0.28 -0.04 -7.7 22 0.61 0.550

90-03-04 decomp./frag. 1.1 2.23±0.23 2.07±0.17 0.16 1.4 22 3.25 0.004

90-03-05 rounded 1.0 7.70±0.76 7.22±0.64 0.48 2.3 12 1.54 0.153

90-03-05 rounded 1.1 8.04±1.37 7.20±0.67 0.83 1.9 11 1.74 0.112

90-03-05 rounded 1.1 8.04±1.37 7.65±0.78 0.39 2.5 11 1.30 0.222

90-03-05 rounded 1.0 2.44±0.37 2.52±0.29 -0.08 -4.5 14 0.83 0.420

90-03-05 rounded 0.9 2.44±0.37 2.52±0.47 -0.08 -4.0 14 0.93 0.368

90-03-05 rounded 0.9 2.52±0.47 2.52±0.29 0.00 134.1 14 0.03 0.978

90-03-06 precip. part. 0.6 0.18±0.04 0.19±0.04 -0.01 -7.3 16 0.55 0.591

90-03-06 precip. part. 0.5 0.18±0.04 0.18±0.04 0.00 10.7 16 0.38 0.713

90-03-06 precip. part. 0.4 0.19±0.04 0.18±0.04 0.01 5.1 16 0.78 0.446

90-03-27 rounded 0.6 2.24±0.45 2.24±0.50 0.00 370 31 0.02 0.988

91-03-30 decomp./frag. < 1 0.69±0.07 0.67±0.09 0.01 7.0 31 0.8 0.430

91-03-30 decomp./frag. < 1 2.03±0.31 2.05±0.20 -0.03 -10.8 27 0.48 0.633

the tapping was discontinued, no significant operator effects (p > 0.05) were detected with

this operator. Since the one experiment with p < 0.01 can be explained and marginally

significant operator effects (0.01 < p < 0.05) only occurred in 2 of 19 experiments,

different operators are not considered to be a source of variability for the relatively small

sets of 7-12 shear frame tests used in subsequent chapters, since the operators received

training and supervision in the early stages of their work.

4.10 Size Effects

Mean shear strengths measured with the shear frame decrease with increasing

frame size (Perla, 1977; Sommerfeld, 1980; Föhn 1987a). This was verified at Mt. St.

Anne in the Cariboo Mountains, where comparisons were conducted using shear frames

with areas of 0.01, 0.025 and 0.05 m2. In each comparison, shear tests with the standard

frame (area 0.025 m2) were alternated with tests with the non-standard frame (area 0.01 or

0.05 m2). Mean differences between adjacent tests, , and the coefficient of variation ofD

the difference, VD, are reported in Table 4.9. Except for the comparison on 1993-03-16,

the larger frame had a lower mean strength than the smaller frame.

The hypothesis that there is no difference in strength is assessed with a t-test

(Eq. 4.7). The probability that there is no difference between mean strengths measured

with the standard and non-standard frames, based on a two-tailed t-test, is p. For 7 of 9

comparisons, the larger frame had a significantly lower mean strength than the standard

frame (p < 0.05). On 1990-04-04, there was no significant difference in mean strengths

(p = 0.41). There was one unexpected and unexplained result. On 1993-03-16, the 0.05 m2

frame had a significantly increased strength compared to the 0.025 m2 frame (p = 2x10-4).

Nevertheless, for eight of nine comparisons the larger frame had a lower mean strength

and for seven of these, the difference was significant.

For strength tests of brittle materials, variability is expected to decrease with

increasing sample size. However, Schaerer (1991) did not find the difference in variability

among frames with areas of 0.01, 0.025 and 0.05 m2 to be significant. The data in Table

4.9 are used to assess the effect of frame area on variability with the F test:

81

F = (stest/sstd)2 (4.8)

where s is the standard deviation for a sample of tests with a particular frame.

For six of the nine comparisons summarized in Table 4.9, variability decreased

with increased frame size. However this decrease was only significant for two

comparisons (p < 0.03). On 1991-12-21, the 0.05 m2 frame showed a significant increase

in variability compared to the 0.025 m2 frame. It appears comparisons involving a greater

number of alternating tests would be required to determine if the effect of frame size on

variability is significant.

To eliminate the effect of frame size from shear strength measurements and relate

shear frame results to much larger areas relevant for slab failure Sommerfeld (1973, 1980)

and Sommerfeld and King (1979) proposed that shear strength could be corrected based

82

Table 4.9 Effect of Shear Frame Area on Mean Strength and Variance

Date Std. Frame TestFrame No.

ofPairs

Difference Variance

MeanStrength

(kPa)

Coef.of

Var.

Area

(m2)

Coef.of

Var.

Mean

(kPa)

Coef.of

Var.

t-test F-test

t p F p

90-02-15 0.691 0.17 0.01 0.12 39 0.23 0.7 8.86 9E-11 0.513 0.979

90-02-25 0.791 0.14 0.01 0.24 32 0.12 1.74 3.2 0.003 2.885 0.002

90-03-17 2.111 0.19 0.01 0.18 54 0.45 1.17 6.22 8E-08 0.875 0.686

90-04-04 3.082 0.16 0.01 0.24 24 0.14 5.72 0.84 0.41 2.218 0.031

90-04-06 3.852 0.13 0.01 0.16 26 0.30 2.30 2.18 0.039 1.526 0.149

91-12-21 0.531 0.08 0.05 0.13 32 -0.08 -0.94 -5.92 2E-06 2.493 0.007

92-02-17 0.642 0.13 0.05 0.11 30 -0.15 -0.46 -11.6 2E-12 0.704 0.825

92-04-10 0.613 0.2 0.05 0.14 32 -0.08 -1.51 -3.70 8E-04 0.52 0.963

93-03-16 2.164 0.17 0.05 0.17 31 0.33 1.31 4.19 2E-04 0.922 0.587

93-04-01 2.214 0.18 0.05 0.16 30 -0.24 -1.71 -3.16 0.004 0.815 0.7081 decomposed and fragmented precipitation particles2 surface hoar3 precipitation particles4 faceted crystals

on Daniels (1945) statistics. Based on a compilation of field studies (Föhn, 1987a), the

Daniels strength, Σ∞, for any loaded area larger than about 0.5 m2, is

Σ∞ = 0. 65 Σ250 (4.9)

and

Σ∞ = 0.56 Σ100 (4.10)

where Σ250, and Σ100 are the shear strengths measured with a 0.025 m2 and 0.01 m2 shear

frame respectively.

For the studies of persistent instabilities reported in subsequent chapters, a

0.025 m2 frame was used for almost all shear tests. On occasions when a force gauge with

sufficient capacity was not available, a 0.01 m2 frame was used. The Daniels strength is

used for all results presented in subsequent chapters.

4.11 Effect of Normal Load

For granular materials, shear strength generally increases with normal load, and for

failures due to yielding, this effect is commonly modelled by the Mohr-Coloumb failure

criterion (de Montmollin, 1982). For brittle fractures, the effect can be empirically

modelled in terms of shear strength, Σ, and normal load, sZZ (Roch, 1966b)

Σφ = Σ + σzzφ(Σ,σzz) (4.11)

where φ(Σ,szz) is the normal load adjustment.

By placing weights on top of shear frames, Roch (1966b) and Perla and Beck

(1983) observed increased strength with increased normal load (Figure 4.6). Roch (1966b)

determined empirical equations for the normal load adjustment for fresh snow, rounded

grains and faceted grains (Eq. 2.3a, b, c).

The previous studies did not provide any results for the effect of normal load on

surface hoar which is very important to avalanche forecasting in western Canada. During

the winter of 1995, the shear strength was measured for three persistent layers and one

non-persistent layer (Table 4.10). For each layer, 6-14 shear frame tests were made in a

83

level study plot with no added normal load and with weights of mass 0.3, 1 and 3 kg

placed on top of a standard shear frame.

The resulting mean Daniels strengths are shown in Table 4.10 and plotted against

the normal stress in Figure 4.7. The normal stress is calculated from the added weights and

does not include the mass of the frame or the snow in the frame which typically total

84

Table 4.10 Effect of Normal Load on the Daniels Strength

NormalStress

σΖΖ(kPa)

Rounded Facets Surface Hoar Rounded Surface Hoar

Decomposed andFragmented

Particles

N DanielsStrength

Mean ± S.E.(kPa)

N DanielsStrength

Mean ± S.E.(kPa)

N DanielsStrength

Mean ± S.E.(kPa)

N DanielsStrength

Mean ± S.E.(kPa)

0 6 0.61 ± 0.02 7 2.31 ± 0.12 8 3.52 ± 0.11 14 1.78 ± 0.06

0.12 6 0.62 ± 0.04 7 2.51 ± 0.06 8 3.49 ± 0.13 0 -

0.39 6 0.68 ± 0.03 7 2.55 ± 0.08 8 3.46 ± 0.10 0 -

1.18 6 0.70 ± 0.07 7 2.63 ± 0.08 8 3.65 ± 0.15 14 2.33 ± 0.03

Figure 4.6 Effect of normal load on strength from previous studies. The ordinate showsthe Daniels strength since Roch (1966a) used a 0.01 m2 frame and Perla and Beck(1983) used a 0.025 m2 frame.

0.3-0.4 kg. For the three persistent layers, the increase in strength is comparable to the

standard errors, and the correlation between the increase in strength and the normal load is

not significant (N = 9, r = 0.43, p = 0.24). This is in contrast to the 0.55 kPa (31%)

increase in strength of the decomposed and fragmented particles.

Since the increase in strength for an increase in normal load is not significant for

persistent layers, no adjustment for normal load (φ = 0) is applied to persistent layers in

subsequent chapters. Further studies are required to determine if there is a small but

significant effect of normal load on the strength of persistent weak layers.

For non-persistent layers, stability indices are calculated in subsequent chapters

using the following equations taken from Roch (1966a) and adjusted to Daniels strength.

For precipitation particles

φ(Σ∞,σzz) = 0.08 Σ∞ + 0.056 + 0.022σzz (4.12)

and for decomposed and fragmented precipitation particles as well as for rounded grains

φ(Σ∞,σzz) = 0.08 Σ∞ + 0.224 (4.13)

85

Figure 4.7 Measured and predicted effect of normal load onDaniels strength.

As shown in Figure 4.7, Equation 4.13 increases too quickly for the persistent

weak layers summarized in Table 4.10 but provides an acceptable prediction for the

normal load effect on the strength of the layer of decomposed and fragmented particles.

4.12 Frame Design

To distribute the applied stress more evenly through the snow layer being tested,

Roch’s (1966a, b) frames had two intermediate cross-members (fins). The relatively rigid

outer frame distributes the manually applied load equally onto the rear cross-member and

the two intermediate cross-members. The lower tip of each of these active cross- members

creates a shear stress concentration in the weak snow layer. These stress concentrations

are influenced by the ratio of the height of the cross-member, d, to the length of the snow

sub-specimen in front of the cross-member, w (Figure 4.8). Perla and Beck (1983)

86

Figure 4.8 Shear frames used for comparative studies of frame design and size effects.

suggested that by decreasing the d/w ratio, stress concentration at the lower tip of the

cross-member is increased, and that increasing d/w will increase the normal load and may

contribute to increased disturbance of the weak layer when the frame is inserted .

Roch’s (1966a) 0.01 m2 frame had 3 active cross-members and a h/w ratio of 3:4.

Perla and Beck (1983) and Sommerfeld (1984) preferred frames with an area of 0.025 m2

but retained the three active cross-members and the slightly trapezoidal shape designed to

minimize friction between the frame and the snow on either side. Perla and Beck (1983)

maintained the d/w ratio of 3:4 (Table 4.11).

Using 0.025 m2 frames, the following effects were studied by alternating tests with

standard and non-standard frames on the same layer:

distance between cross-members (w = 31 mm compared with standard w = 52 mm)

reduced frame height (d = 20 mm compared with standard d = 40)

87

Table 4.11 Shear Frame Specifications

FrameIdentifier

Area(m2)

No. ofActiveCross-

Members

Cross-MemberHeight

d(mm)

Dist.BetweenCross-

Membersw

(mm)

FrameWidth

toLength Ratio

MaterialThickness

(mm)

FrameMaterial

Massof

Frame

(kg)

standard 0.025 3 40 52 1:1 0.6 st. steel 0.20

short 0.025 3 20 52 1:1 0.6 st. steel 0.12

5-fin1 0.025 5 40 31 1:1 0.6 st. steel 0.30

100 0.01 3 25 35 1:1 0.8 st. steel 0.09

500 0.05 4 40 53 1:1 0.8 st. steel 0.48

Swiss 0.05 5 30 47 4:5 1.5 st. steel 0.87

finger-fin 0.025 32 30 18 1:1 0.6 st. steel 0.301 Also referred to as the 5-cross-member frame

4.12.1 Reduced Distance Between Cross-Members

The effect of reduced distance between fins was studied by alternating tests with a

5-cross-member frame (w = 31) with tests using a standard frame (w = 52 mm, 3 active

cross-members). In each of the five comparisons (Table 4.12), the mean strength

measurement was reduced by 16-26% by decreasing the distance between cross-members

and the difference was significant (p < 10-6 ). Similarly, Perla and Beck (1983) reported a

15% reduction in strength measurement when the distance between cross-members was

reduced from approximately 50 mm (3 active cross-members) to approximately 30 mm (5

active cross-members). Increasing the number of cross-members increases the number of

stress concentrations resulting in reduced strength measurements.

4.12.2 Reduced Frame Height

The effect of reduced frame height (d = 20 mm) was compared with standard

height frames (d = 40 mm) by alternating tests on particular weak layers. The resulting

differences in mean strength measurement are both positive and negative. Only after

considering whether the frame was placed in or above the weak layer, did a pattern

become apparent (Figure 4.8). For the eight comparisons in which both the standard and

the short frame were placed in the weak layer, t-tests showed the mean strength

measurements to be not significantly different in four comparisons, and that the mean

strength measurement with the short frame was significantly greater than the mean

strength measurement with the standard frame in the other four comparisons (Figure 4.9).

Greater strength measurement for the short frame when both frames are placed in the

weak layers is difficult to explain.

When both the standard and short frame were placed above the weak layer, the

mean strength measurement with the short frame was less than the mean strength

measurement with the standard frame in each of the four comparisons and the difference

was significant for three of the four comparisons. The reduced strength measurement with

the short frame is probably due to increased stress concentrations associated with the

shorter cross-members (Perla and Beck, 1983).

Two points are important to interpret these potentially confusing results:

88

89

Table 4.12 Effect of Shear Frame Design on Mean Strength

Date MicrostructureStd. Frame Test Frame Difference t-test

MeanStr.

(kPa)

C. of

Var.

I.D.(Table4.11)

C. of

Var.

Mean

(kPa)

C. of Var.

No.of

Pairst p

92-02-14 surface hoar 0.39 0.08 5-fin 0.14 -0.07 -0.78 30 7.02 1.0E-07

93-02-13 surface hoar1 3.18 0.14 5-fin 0.18 -0.56 -1.04 32 5.45 6.0E-06

93-02-24 surface hoar 0.48 0.14 5-fin 0.15 -0.08 -0.96 32 5.88 1.7E-06

95-02-22 surface hoar,facets

1.87 0.12 5-fin 0.09 -0.29 -0.84 30 6.56 3.5E-07

95-03-07 surface hoar1 4.18 0.08 5-fin 0.16 -1.09 -0.58 30 9.48 2.2E-10

91-12-20 precip. particles2 0.34 0.10 short 0.11 0 -38.4 36 0.16 0.88

93-02-24 surface hoar2 0.48 0.14 short 0.13 0.02 3.94 32 1.44 0.16

93-03-03 surface hoar,facets2

0.61 0.15 short 0.16 0.11 1.31 30 4.19 2.4E-04

94-03-30 graupel2 4.01 0.12 short 0.18 0.87 1.04 30 5.27 1.2E-05

94-12-04 graupel2 0.93 0.14 short 0.11 0.02 8.09 30 0.68 0.50

95-01-04 graupel2 2.12 0.16 short 0.19 -0.09 -5.19 30 1.06 0.30

95-02-09 surface hoar 3.71 0.09 short 0.14 -0.09 -6.55 33 0.88 0.39

95-03-01 surface hoar 3.3 0.10 short 0.1 -0.24 -1.65 31 3.36 2.1E-03

95-03-24 graupel2 2.76 0.15 short 0.14 0.38 1.54 30 3.55 1.3E-03

95-03-28 graupel2 3.6 0.10 short 0.13 0.34 1.56 30 3.51 1.5E-03

95-03-29 surface hoar 5.92 0.09 short 0.13 -0.19 -3.33 30 1.64 0.11

95-03-17 surface hoar1 3.29 0.11 short 0.12 -0.3 -1.69 26 3.01 0.01

92-03-27 precip. particles 0.31 0.11 Swiss 0.12 0.13 0.59 14 6.3 2.7E-05

92-04-10 facets 0.61 0.2 Swiss 0.24 0.39 0.53 32 10.6 8.1E-12

94-03-30 graupel2 4.01 0.12 Swiss 0.21 0.42 2.49 30 2.2 0.04

95-02-28 surface hoar 3.03 0.08 finger-fin

0.11 0.49 0.83 36 7.26 1.8E-08

95-03-18 graupel2 1.55 0.10 finger-fin

0.09 0.88 0.37 10 8.64 1.2E-05

1 rounding or surface hoar crystals apparent2 bottom of frame placed in weak layer

Frames can usually be placed above the weak layer for the persistent weak layers

that are so important to avalanche forecasting. For such frame placements, the short

frame resulted in strength reductions of only 2% to 9%.

The difference in mean strength measurement was only significant for one of five

comparisons for mean strengths below 2.6 kPa. This lower range is particularly

relevant to avalanche forecasting since the mean strength for the critically weak

layers that released 38 of the 40 skier-triggered slabs tested with a standard frame

(Chapter 7) was less than 1.6 kPa.

Also, the front cross-member (that does not directly apply load to the specimen) of

the short frame was observed to bend frequently when testing stronger layers. A thicker,

stiffer front cross-member would reduce bending but could contribute to weak layer

disturbance during frame placement. This bending problem combined with the fact that

operators found the short frame difficult to place with respect to the weak layer question

the merit of further studies with the short frame.

90

Figure 4.9 Twelve strength comparisons of short frame with standard frame.Significance levels less than 0.1 are shown.

4.12.3 The Swiss Shear Frame

The Swiss shear frame used by Föhn (1987a) differs in every dimension and

proportion from the others listed in Table 4.11. Most noticeably, the sheet metal is 2.5

times the thickness of the standard frame and the weight more than four times that of the

standard frame.

In three comparisons in which tests with the 0.05 m2 Swiss frame were alternated

with the 0.025 m2 standard frame, the mean strength measurements obtained with the

Swiss frame were 10%, 42% and 63% greater than the strength measurements obtained

with the standard frame (Table 4.12). Such increases cannot be explained simply by the

increase in normal load due to the mass of the Swiss frame since such increases would

range between 1 % and 14% according to Eq. 4.12.

The Swiss frame requires considerably more insertion force because it has more

cross-members of thicker metal than the standard frame. The disturbance due to pushing

the thicker cross-members through the snow and close to the weak layer could cause

additional bonding and perhaps a measurable strength increase due to “fast

metamorphism” (Gubler, 1982; de Montmollin, 1982) caused by the additional pressure on

the weak layer.

4.12.4 The Finger-Fin Shear Frame

Lang and others (1985) used a very different shear frame for a study of surface

hoar. The 0.01 m2 frame was designed by R.L. Brown and R. Oakberg to reduce stress

concentrations by eliminating the active cross-members and using 32 “finger-fins”, each

10 mm wide, extending down from a top plate. Further, the fins were 8 mm shorter than

the side walls to ensure that the fins did not penetrate into the weak layer.

For comparison with the standard 0.025 m2 frame, a 0.025 m2 version of Brown

and Oakberg’s finger-fin frame was built (Figure 4.6). Each of the 32 fins were 17 mm

wide, 30 mm long and were 10 mm shorter than the sides of the frame. On 28 February

1995 and 18 March 1995, the frame resulted in higher mean strength measurements

(Table 4.12), presumably due to the reduced stress concentrations. However, on 18 March

91

1995, it took 30 attempts to get 10 fractures on the weak layer being tested. The other 20

tests were rejected because the fracture occurred near the bottom of the finger-fins in the

snow above the weak layer. There was no problem with the standard frame. A similar

problem occurred on 24 March 1995, when no fractures occurred in the weak layer being

tested during 14 attempts. Again, the standard frame produced consistent planar fractures.

As discussed in Section 4.8, snowpack conditions often dictate that the shear

frame be placed a certain distance above the weak layer, usually 0-10 mm. Since the

finger-fin frames are designed to locate the fin-tips 8-10 mm above the weak layer, they

cannot test as many weak layers as the standard compartmental frame.

4.13 Summary

Shear strength measurements from shear frame tests are assumed to be normally

distributed since only 4 to 8 of 28 sets of 30 or more tests show evidence of

non-normality (Section 4.2).

Coefficients of variation for shear frame tests average 0.15 and 0.18 from level

study plots and avalanche start zones respectively (Section 4.3). These values are

less than the 0.25 reported in previous studies and result in a reduced number of

tests to achieve a particular level of precision.

Shear frame tests that result in divots more than 10 mm deep under the rear

compartment of the frame yield strength measurements significantly greater than

tests with planar fractures (Section 4.4). No significant effect could be detected for

10 other common shapes of fracture surfaces.

Shear frame strengths tend to increase at slower loading rates. However, the effect

of loading rate on strength is reduced for loading times less than 1 second, and is

negligible for mean strengths less than 1 kPa (Section 4.5). This is consistent with a

laboratory study (Fukuzawa and Narita, 1993) using constant displacement rates

that found brittle fractures and a reduction in strength of only 12% when loading

times were reduced from 1.5 to 0.2 seconds.

92

The first two tests in a set of tests are more variable than subsequent tests and could

be rejected to improve within-set variability (Section 4.6).

Although not recommended, delays of up to 3 minutes between placing the frame

and pulling the frame do not appear to affect the resulting shear strength

measurements (Section 4.7).

Placing the bottom of the frame in the weak layer results in lower strengths than

placing the bottom of the frame a few mm above the weak layer (Section 4.8).

Frame placements 2-5 mm above the weak layer usually result in planar fractures,

but frames must sometimes be placed in the weak layer or more than 5 mm above it

to obtain planar (shear) failures.

With consistent technique, there is no apparent difference in mean strength

measurements obtained by different experienced shear frame operators (Section 4.9)

using the same approximate loading rate and technique for placing the frame.

Shear frames with larger areas result in lower mean strengths than smaller frames

(Section 4.10) as shown in previous studies. Although strength measurements

obtained with larger frames usually show reduced variance compared to smaller

frames, the reduction is not statistically significant. Based on the work of

Sommerfeld (1980) and (Föhn, 1987a), strength measurements obtained with

0.01 m2 and 0.025 m2 frames are adjusted to the equivalent strength of a very large

specimen, called the Daniels strength. In subsequent chapters, the strength

measurements obtained with the 0.025 m2 standard frame are multiplied by the

appropriate adjustment factor, 0.65, to obtain the Daniels strength.

Persistent weak layers of surface hoar and rounded facets do not show a significant

strength increase with increased normal load. This is in contrast to the increase

reported by Roch (1966b) for depth hoar and for non-persistent microstructures

(Section 4.11). No adjustment for normal load (φ = 0) is applied to the strength of

persistent weak layers in subsequent chapters. If a weak normal load effect exists for

persistent layers—and more extensive field studies are recommmended—then the

93

stability indices in subsequent chapters for thick, dense slabs overlying persistent

weak layers may be conservative.

Decreasing the distance between active cross-members while keeping the overall

dimensions of the frame constant increases the number of stress concentrations and

reduces the mean shear strength measurement (Section 4.12).

Decreasing the height of the active cross-members has an inconsistent effect, tending

to increase the strength measurement when the bottom of the frame is placed in the

weak layer and decrease the strength measurement when the bottom of the frame is

placed above the weak layer. However, this study was complicated by bending of

the front cross-member of the shorter frame and by difficulty placing the shorter

frame.

The relatively heavy Swiss shear frame results in increased shear strength

measurements compared to the 0.025 m2 shear frame used as a standard in the

present study (Section 4.12.3).

The finger-fin shear frame results in decreased shear strength measurements due to

reduced stress concentrations but restricts the operator's ability to place the frame a

certain distance above the weak layer (Section 4.12.4), a practice that is often

required to obtain planar fractures in particular weak layers.

94

5 FINITE ELEMENT STUDIES OF THE SHEARFRAME TEST

5.1 Introduction

Although sloping snowpacks have been studied with finite element models (Curtis

and Smith, 1975; Singh, 1980; Bader and others, 1989; Bader and Salm, 1990; Schweizer,

1993), a literature review revealed no finite element models of the shear frame test. In

fact, neither analytical nor finite element models for the stress distribution have dealt with

the effect of cross-member height or spacing between cross-members proposed by Perla

and Beck (1983). In this chapter, a simple finite element model is developed and used to

qualitatively assess the effect on the shear stress distribution due to:

the stiffness of the snow within the frame,

the placement of the shear frame with respect to the weak layer, and

cross-member height and spacing between cross-members.

5.2 The Model and Assumptions

The basic geometry of the two dimensional model is shown in Figure 5.1. The

models consist of three isotropic layers: the snow within the frame (superstratum), the

weak layer and the bed surface (substratum). The superstratum is modelled as either part

Figure 5.1 Geometry and loading for finite element model of standard shear frameplaced 3 mm above weak layer.

95

of a soft slab (~200 kg/m3) or part of a hard slab (~400 kg/m3). The weak layer is

modelled as a 2-mm thick softer layer (~160 kg/m3). Material properties for the three

layers were chosen from Mellor’s (1975) compilation of snow properties and are shown in

Table 5.1. In four of the six models summarized in Table 5.2, the bottom of the frame is

3 mm above (ab) the weak layer. In the remaining two models, the bottom of the frame is

1 mm into the 2-mm thick weak layer.

Table 5.1 Material Properties for Finite Element Model

Layer Nominal Density(kg/m3)

Young’s Modulus(MPa)

Poisson’s Ratio

Soft Superstratum 200 10 0.25

Hard Superstratum 400 100 0.25

Weak Layer 160 2 0.25

Bed Surface 200 10 0.25

Table 5.2 Finite Element Models of the Shear Frame Test

ModelName

ShearFrame

FramePlacement with

Respect toWeak Layer

Stiffness ofSuperstratum

E (MPa)

No. ofElements

Displacement Necessaryfor Average Shear

Stress of 1 kPa (mm)

std-ab-soft std 3 mm above 10 7 130 0.0130

std-in-soft std 1 mm into 10 6 604 0.0130

std-ab-hard std 3 mm above 100 7 130 0.0077

5-ab-soft 5-fin1 3 mm above 10 7 110 0.0115

shrt-ab-soft short 3 mm above 10 5 570 0.0136

shrt-in-soft short 1 mm into 10 5 044 0.01361 frame has 5 cross-members or fins

The bed surface is fixed (0 displacement) 30 mm to the right of the frame, 20 mm to

the left of the shear frame and at the base 30 mm below the weak layer. The left surfaces

of the snow within the frame compartments were loaded with constant displacement to the

right, in preference to pressure loading which would have tended to tilt the snow in the

compartments unrealistically. The constant displacement for the left surface of each

96

compartment is a consequence of assuming that the frame is rigid. The displacement was

chosen to cause an average shear stress in the weak layer of 1.0 kPa which is typical of the

shear strength of persistent weak layers (Table 5.2).

A linear model is used since shear frame loading times (< 1 s) are well within the

range associated with linear stress-strain curves and brittle failures for tension (Narita,

1980, 1983; Singh, 1980) and for shear (Fukuzawa and Narita, 1993). However, such

macroscopic linear behaviour does not rule out small-scale plasticity at stress

concentrations and grain boundaries. Nevertheless, linear elasticity is assumed since it is

sufficient to provide qualitative comparisons of different frame designs, frame placements

and material properties. Since the sides of the frame restrict expansion during loading, a

two-dimensional plane strain model is used.

Each element is a bi-linear quadrilateral with nodes at each corner and the midpoint

of each of the four sides. Such elements require additional calculations compared to

four-node quadrilateral elements but they define a quadratic shape function which allows

the sides of the elements to curve during deformation. In and near the lower tips of the

active cross-members and the weak layer, the elements are 1 mm by 1 mm prior to loading

as shown in Figure 5.2. At the upper and lower surfaces of the model, well away from the

weak layer, the size of the elements increases to 4 mm by 1 mm to reduce the number of

elements and consequently the number of computations.

Elements are joined at nodes, providing continuity. Boundary conditions, such as

displacements, are applied at the nodes.

As a consequence of the assumed linear elasticity, the peak stresses depend strongly

on the size of the elements—smaller elements resulting in higher peak stresses. For the

following comparisons, the same mesh of elements in and near the weak layer is used for

all models. Thus, the peak stresses reflect—at least relatively—the various geometries and

the material properties that are being compared.

The models, material properties and boundary conditions were encoded using Patran

software. Finite element calculations were done by Abaqus software.

97

5.3 Basic Stress Distribution

The contour plot of σXZ shows a stress concentration at each of the three active

cross-members plus one at the right cross-member (Figure 5.3). This rightmost stress

concentration indicates the effect of cutting around the frame with a blade through the

3 mm of superstratum below the frame and into the weak layer. The stress concentration

at the leftmost cross-member is partly due to the applied displacement and partly due to

the blade notching the weak layer. The superposition of these two stress concentrations

results in the peak stress near the leftmost (back) cross-member (Figure 5.4). This explains

why fracture surfaces with divots under the left (back) compartment (Table 4.3) are much

more common than divots under the middle or right (front) compartments. The stress

concentration at the three active cross-members is two-lobed. These two lobes extend into

the weak layer which is 3 mm below the bottom of the cross-members and can be seen as

peaks in σXZ as shown in Figure 5.4.

Figure 5.2 Finite element mesh for snow in left compartment andunderlying weak layer and substratum.

98

Figure 5.3 Stress contours for σxz for standard frame placed in soft superstratum 3 mmabove weak layer. Displacement is 0.013 mm to the right on the left edges of the threecompartments of snow in the frame.

Figure 5.4 Shear stress σXZ in weak layer for standard frame placed 3 mm above weaklayer representing average σXZ of 1.0 kPa.

99

5.4 Effect of Frame Placement on Stress Distribution

The distance between the weak layer and the bottom of the shear frame as

recommended in Chapter 3 is 2-5 mm. However, in practice the distance necessary to

achieve planar shear failures ranges between 0 and 20 mm depending the strength of the

weak layer and the hardness of the layer above the weak layer. The distribution of σXZ in

the weak layer is modelled for two common frame placements: frame placed 1 mm into a

2-mm thick weak layer and 3 mm above a 2-mm thick weak layer. As shown in Figure 5.5,

the peak values of σXZ are reduced when the frame is placed 3 mm above the weak layer

compared to frame placements into the weak layer. This is consistent with field studies

(Section 4.8) in which frames placed in weak layers resulted in lower strengths than

frames placed 2-10 mm above weak layers. The analysis confirms that the more even

stress distribution when the frame is placed above the weak layer is advantageous and

should be used whenever practical.

Figure 5.5 Shear stress σXZ in weak layer for standard frame placed in weak layer and3 mm above weak layer.

100

5.5 Effect of Frames Placed in Hard and Soft Slabs

Ideally, a strength test should depend only on the mechanical properties of the test

specimen. However, practical strength tests fall short of this ideal. The shear strength

measured with the shear frame test depends on factors such as the frame design, the

loading rate and the distance between the frame and the weak layer.

The stiffness of the snow gripped by the frame (superstratum) may also affect the

distribution of shear stress of the weak layer and strength. Since the same weak layer can

not be tested in situ with superstrata of varying stiffness, finite element models are used.

Young’s Moduli of 10 MPa and 100 MPa are chosen to represent soft (~200 kg/m3) and

hard slabs (~400 kg/m3). For both models, the shear frame is 3 mm above a 2-mm thick

weak layer. As shown in Figure 5.6, the distribution of σXZ within the weak layer is more

even for the hard slab than for the weak slab. The stiffer slab results in reduced stress

peaks which would tend to increase the measured strength of the weak layer. Although the

difference in stress distributions due to the varied stiffness of the superstratum is much less

than the difference caused by placing the shear frame above or into the weak layer

101

Figure 5.6 Distribution of σXZ for the standard frame placed in soft and hard superstrata.In both cases, the frame is 3 mm above the weak layer.

(Figure 5.6), the stiffness of the snow gripped by the frame is one more factor that can

affect the measured strength of a weak layer.

5.6 Effect of Spacing Between Cross-members

In the field study summarized in Section 4.12, increasing the number of active

cross-members from three to five while keeping the other dimensions of the frame

constant consistently reduced the measured strength of the weak layer. To compare with

that field study, a shear frame test with a 5-cross-member frame was also modelled with

finite elements. For standard and 5-cross-member models (Figure 5.7), the slab was soft

(E = 10 MPa) and the frames were 3 mm above the 2-mm thick weak layer. The

distribution of σXZ for both the standard frame and the 5-cross-member frame are shown in

Figure 5.7. For the 5-cross-member frame, σXZ has two more peaks, but all peaks are

reduced in magnitude compared with those caused by the standard frame. Based on the

assumption that the strength is determined by the peak stress, the reduced peaks

102

Figure 5.7 Distribution of σXZ for 5-cross-member and standard frame.

associated with the 5-cross-member frame would suggest increased strength rather than

decreased strength as measured in the field. There are at least two possible explanations:

The assumption of strength being determined by peak stress is too simplistic.

Specifically, the linear elastic model ignores plasticity in the snow near the lower

edges of the cross-members.

Although the peak stresses are reduced with the 5-cross-member frame, the number

of peaks is increased by two. Thus the probability of a stress concentration due to a

cross-member being near a flaw in the weak layer is increased.

A more detailed model based on elasto-plasticity, or a non-continuum model based

on a probabilistic distribution of bonds and “chains” (Kry, 1975; Gubler, 1978) is beyond

the scope of this study.

5.7 Effect of Cross-Member Height

The effect of cross-member height was assessed by comparing the distribution of σXZ

within the weak layer for the standard frame (40 mm cross-members) with the short frame

(20 mm cross-members). The shear stress distribution is plotted in Figure 5.8 for both

frames placed 3 mm above the 2 mm thick weak layer, and in Figure 5.9 for both frames

placed 1 mm into the 2 mm-thick weak layer. In both cases, the difference between the

stresses induced by the standard and short frame is minimal. This is consistent with the

field studies (Section 4.12.2) which only detected a significant difference for 1 of the 5

comparisons with mean strengths less than 2.5 kPa. However, it is in contrast with Perla

and Beck (1983) who proposed that reducing the frame height while keeping the distance

between cross-members constant would concentrate the shear stress closer to the

cross-members. This effect is not apparent in Figure 5.8 or 5.9.

For particular material properties and geometry of the model, σXZ is proportional to

the displacement due to the assumption of linear behaviour. Thus, for average shear

strength values in the 2.5 to 6 kPa range in which the in situ comparisons (Section 4.12.2)

detected a significant effect of cross-member height, the model would also predict the

same stress distribution for the short and standard height frames. Similarly, if the Young’s

103

104

Figure 5.9 Distribution of σXZ in weak layer for standard and short frames placed 1 mminto the weak layer. The line for short frame is shifted 5 mm to the left for clarity.

Figure 5.8 Distribution of σXZ in weak layer for standard and short frames placed 3 mmabove the weak layer. The line for short frame is shifted 5 mm to the left for clarity.

Moduli for the superstratum, weak layer and substratum are scaled while maintaining the

10:2:10 ratio (Table 5.1), then σXZ will also be scaled by the same factor for the models

with short and standard frame heights. Hence, the finite element model does not show any

substantial effect of frame height on σXZ. This is consistent with the explanation offered in

Section 4.12.2 that the measured difference between short and standard frames for

relatively strong weak layers may be due to increased bending of the cross-members in

the short frame.

5.8 Summary

There are stress concentrations associated with the active cross-members and with

the notching of the weak layer caused by cutting along the front and back of the

frame with a blade. However, such cutting is essential to ensure that a specimen of

known size, free from restraint by the adjacent snowpack, is tested.

The stiffness of the snow within the frame influences stress concentrations and

consequently the measured strength of the weak layer a few mm below the frame,

although other factors such as the distance between the bottom of the frame and the

weak layer may have a greater effect on stress concentrations, and consequently, on

measured strength.

Placing the frame 2-5 mm above the weak layer reduces stress concentration and is

recommended whenever practical.

The finite element model for the 5-cross-member frame shows two additional stress

concentrations compared to the model for the standard 3-cross-member frame.

However, the peak stresses for the 5-cross-member frame are reduced compared to

the standard frame. Since the measured strength with the 5-cross-member frame is

not increased compared to the standard 3-cross-member frame (Section 4.12.1),

either strength is not determined simply by peak stress or the assumption of linear

elasticity is too simplistic to model the stress concentrations at the lower edges of

the cross-members.

105

According to the finite element models for frames placed 3 mm above the weak

layer and for frames placed 1 mm into a 2-mm thick weak layer, the distribution of

shear stress within the weak layer is not substantially affected by reducing

cross-member height from 40 mm to 20 mm.

106

6 SHEAR FRAME RESULTS AND STABILITYINDICES

6.1 Introduction

While the emphasis in this chapter is on relating stability indices to natural and

skier-triggered dry slab avalanches, Sections 6.2 and 6.3 relate the shear strength of weak

layers to density and hand hardness.

Before assessing a stability index for natural avalanches based on detailed on-site

investigations by researchers, this limited set of investigated avalanches is shown to be

similar to a much larger but less detailed set of avalanches reported by ski guides for the

Columbia Mountains (Section 6.4).

Shear frame stability indices, SN and SN38 are assessed for natural avalanche activity

on test slopes and in surrounding terrain in Sections 6.5 and 6.6, respectively. Shear frame

stability index SS is related to skier-triggered slab avalanche activity on test slopes in

Section 6.7 and refined to obtain the stability index SK in Section 6.8. An “extrapolated”

variation of SK called SK38 is assessed for skier-triggered slab avalanche activity in Section

6.9.

6.2 Shear Strength of Weak Layers Related to Density

The shear strength of dry snow is strongly related to density (e.g. Keeler and Weeks,

1968; Keeler, 1969; Mellor, 1975; Perla and others, 1982) and microstructure (e.g. Keeler

and Weeks, 1968, Perla and others, 1982; Föhn, 1993). While laboratory studies have

identified a decrease in the tensile strength of dry snow with an increase in temperature

(Roch, 1966b; Narita, 1983), such an effect has proven difficult to identify in field studies

of shear strength (Perla and others, 1982) or tensile strength (Jamieson, 1989). Although

liquid water content also affects strength (e.g. Brun and Rey, 1987), it is not a factor in

the present study which is restricted to dry snow. Grain size is not considered a predictor

of strength since previous field studies have not established a significant effect (Perla and

others, 1982; Jamieson, 1989; Föhn, 1993).

107

Perla and others (1982) reported shear strength of weak layers as a function of

density but many density samples included snow from adjacent layers since the weak

layers where thinner than their density sampler (20 mm). Föhn (1993) reported the shear

strength of weak layers and interfaces but did not relate shear strength to density,

presumably since many of the weak layers were too thin to be sampled for density.

Further, since the failure planes for slab avalanches are often interfaces (Föhn, 1993), it

would seem that density samples of failure planes and hence a relationship between the

shear strength of failure planes and density are impossible. However, for approximately

17% of the failure planes tested with shear frames in the present study, the grains at the

failure plane were indistinguishable from an adjacent layer (superstratum or substratum),

and the adjacent layer was thick enough (> 35 mm) for density sampling. The shear

strength of these weak planes is related to the density of the indistinguishable adjacent

layers in this section. Since surface hoar is always too thin for density sampling and always

distinct from adjacent layers that are thick enough for density sampling, no data for

surface hoar are included in this section.

The dependence of tensile and shear strength on density is non-linear (e.g. Keeler

and Weeks, 1968; Keeler, 1969; Martinelli, 1971; Mellor, 1975; Perla and others, 1982;

Jamieson and Johnston, 1990). Ballard and Feldt’s (1965) theoretical model for sintering

of rounded grains is inappropriate since the microstructure of many of the weak layers in

the present study are not rounded and show little evidence of sintering. Also, Perla and

others (1982) obtained a better fit to shear frame strengths with the relation

Σ = A(ρ/ρice)B

(6.1)

where ρice is the density of ice (917 kg/m3) and A and B are empirical constants that

depend on microstructure.

Since the variance of snow strength increases with the mean strength, Martinelli

(1971) and Jamieson and Johnston (1990) stabilized the variance with a logarithmic

transformation. A log-log transformation of Eq. 6.1 yields

ln Σ = ln A + B ln (ρ/ρice) (6.2)

108

For microstructure classes 1 to 5 (Colbeck and others, 1990), the empirical variables

A and B and the coefficients of determination, R2, are shown in Table 6.1 for regressions

of Daniels strength on density with and without the logarithmic transformation. The mean

Daniels strengths are also plotted in Figure 6.1. Graupel (class 1f) is distinguished from

other types of precipitation particles with a different symbol. As reported by a field study

of tensile strength (Jamieson and Johnston, 1990), layers of graupel are generally weaker

than other types of precipitation particles with the same density. Similarly, Figure 6.1 uses

a different symbol for rounding facets (class 4c) than for other types of facets. As shown

in Table 6.1, the seven mean strengths of rounding facets are not correlated with density.

However, the mean strengths of the rounding facets fall within the 46 mean strengths for

faceted grains, and the subclass is subsequently included within the class of faceted grains.

109

Table 6.1 Strength-Density Regressions by Microstructure

Microstructure(Colbeck andothers, 1990)

No. ofMean

Strengths

Densityρ

(kg/m3)

RegressionΣ∞ = A(ρ/ρice)

B Regression

ln Σ∞ = ln A + B ln (ρ/ρice)

A B R2 A B R2

PrecipitationParticles1 (1)

11 50-110 8.3 1.55 0.44 3.09 1.18 0.33

Graupel (1f) 3 110-235 - - - - - -

Decomposed/Fragmented (2)

65 65-270 24.9 2.07 0.67 12.8 1.74 0.42

Rounded Grains(3)

12 105-270 12.2 1.52 0.57 10.2 1.44 0.60

FacetedCrystals2 (4)

46 110-330 16.4 1.94 0.35 22.3 2.25 0.57

Rounding Facets(4c)

7 205-280 1.35 -0.10 0.00 1.15 -0.20 0.01

Depth Hoar (5) 2 250-280 - - - - - -

Group I 1 (1,2,3) 88 50-270 23.0 2.00 0.73 13.7 1.76 0.63

Group II (4,5) 55 110-330 14.0 1.80 0.32 22.8 2.23 0.541 excluding graupel (1f)2 excluding rounding facets (4c)

In the Columbia Mountains where most of the strength measurements were made,

precipitation particles (class 1) generally metamorphose into decomposed grains (class 2)

which in time metamorphose into rounded grains (class 3). Not surprisingly, these three

microstructures show a continuous increase in strength with increasing density, and are

assembled as Group I microstructures for additional regressions shown in Table 6.1. A

similar trend for Group I microstructures has also been shown for tensile strength

(Jamieson and Johnston, 1990).

Commonly in the Rocky Mountains of western Canada and occasionally in the

Columbia Mountains, faceted grains (class 4) metamorphose into depth hoar (class 5). In

Figure 6.1, these two microstructures show a similar increase in strength with increasing

density. They are assembled into Group II for additional regressions in Table 6.1. For the

regressions of Daniels strength on density using Equations 6.1 and 6.2, the coefficients of

determination for the 55 points with Group II microstructures are 0.32 and 0.54

respectively. For the 88 points with Group I microstructures, the corresponding

110

Figure 6.1 Daniels strength for weak layers by microstructure and density.

coefficients of determination are 0.73 and 0.63. Since the Group II microstructures show

reduced coefficients of determination for fewer points, the mean shear strengths are clearly

more variable as a function of density. This is consistent with previous studies for tensile

strength (Sommerfeld, 1973; Jamieson and Johnston, 1990).

As shown in Table 6.1, the log-log transformation reduces the coefficient of

determination for Group I microstructures, and increases it for Group II microstructures.

Since the intent of the transformation was to stabilize the variance, the preferred

regression will be the one with the more consistent variance. For both regressions, the

variance for five strength intervals is plotted for Group I and II in Figure 6.2. For this

comparison, the variance for each interval is normalized using the total variance for the

entire range of strengths. For both microstructure groups, the log-log transformation

increases the normalized variance for low strengths and decreases it for high strengths.

However, Group I microstructures show a more consistent variance without the

transformation and the Group II microstructures show a more consistent variance with the

transformation. The greatest normalized variance is, not surprisingly, for the highest

strengths and there are more high strengths for Group II microstructures. Martinelli

(1971) used the logarithmic transformation for strengths that ranged up to 100 kPa and

Jamieson (1989) found that the transformation effectively stabilized the variance for tensile

111

Figure 6.2 Normalized regression variance for Group I and II microstructures.

strengths that ranged up to 8 kPa. For the generally low shear strengths associated with

weak layers, the transformation is only effective for the highest measured strengths.

Nevertheless, the preferred regressions are the ones with the most consistent variance (and

highest coefficients of determination). The best fit for Group I microstructures is obtained

with the regression based on Eq. 6.1 and with its logarithmic transformation (Eq. 6.2) for

Group II microstructures.

For the range of densities reported in Table 6.1 and Figure 6.1, the regression line

for Group II microstructures falls below the line for Group I microstructures. This is

consistent with field observations that faceted grains are weaker than partly decomposed

and rounded grains with the same density.

6.2.1 Comparison with Previous Field Study

Perla and others (1982) also reported shear strength as a function of density for the

common microstructures. Although they used a very similar frame (0.025 m2 with three

active cross-members), there are several relevant differences. In the present study:

The shear frames are pulled to failure within 1 s, resulting in brittle fractures

—according to Fukuzawa and Narita (1993)—whereas some of Perla and others

(1982) results could involve ductility since they were pulled to failure “within a few

seconds”.

The strengths are only plotted against density and regressed on density when the

resistance and grain type of the weak plane are indistinguishable from an adjacent

weak layer that is thick enough for a density sampler, whereas Perla and others

(1982) took density samples centred on the weak layer and hence included snow

from the layers above and below the weak layer whenever the weak layer was

thinner than their density sampler.

The data are exclusively for weak layers whereas most of the results from Perla and

others (1982) are for homogeneous layers.

Perla and others (1982) report regression parameters similar to those in Table 6.2

and based on Eq. 6.1. Their regressions for the 0.025 m2 shear frame are readily converted

112

to Daniels strength by multiplying the coefficient A in Eq. 6.1 by 0.65 (Sommerfeld, 1980;

Föhn, 1987a). For the four microstructures common to Perla and others (1982) and the

present study, the regressions on density are compared in Figure 6.3.

As shown in Figure 6.3, Perla and others (1982) report shear strengths over a wider

density range, presumably because they did not restrict their tests to weak layers. Only for

layers of precipitation particles did Perla and others (1982) report lower strengths, and

then only by approximately 0.1 kPa. For decomposed and fragmented precipitation

particles, both studies report similar strengths, although strengths from the previous study

are approximately 0.2 kPa higher for a density of 250 kg/m3. Similarly for rounded grains

and faceted crystals, the strengths from Perla and others (1982) are substantially higher

than those from the present study for densities greater than 250 kg/m3. Such differences

are not surprising since the present study was restricted to weak layers and to loading

times of less than 1 s, both of which are associated with lower strengths.

Figure 6.3 Shear strengths from present study compared with those from Perla andothers (1982) for four common microstructures.

113

6.3 Shear Strength of Weak Layers Related to Hand Hardness

The most widely used measure of resistance in Canada and internationally is “hand

hardness”, which results from a simple, quick and empirical test. A fist, four finger tips,

one finger tip, the blunt end of a pencil or a knife tip is pushed horizontally into a snow

layer while wearing gloves. The hand hardness is simply the bluntest object that can be

pushed into the snow with a force of 10-15 N in Canada (NRCC/CAA, 1989; CAA, 1995)

or 50 N internationally (Colbeck and others, 1990). (For the present study, the hand

hardness of layers as thin as 3 mm were tested using a thin plastic ruler to compare their

resistance with a thicker layer in the same pit for which the hand hardness could be

determined with a fist, fingers, pencil or knife.) The levels of hand hardness are

abbreviated as F, 4F, 1F, P and K. Snow layers with resistance between the five major

levels of hand hardness can be qualified with a “+” or “-” sign, giving 15 levels: F-, F, F+,

4F-, 4F, 4F+, 1F-, 1F, 1F+, P-, P, P+, K-, K and K+ (CAA, 1995). Ice layers harder than

“knife” are labelled I for

ice.

For the four most

common microstructures,

mean Daniels strengths

are plotted for hand

hardness classes of F, 4F,

1F and P in Figure 6.4.

Along with each mean,

“whiskers” show the

range of two standard

errors. An increase in

variability as indicated by

the standard error is

apparent for increasing mean

strength. For sets of more

114

Figure 6.4 Shear strength by hand hardness for commonmicrostructures. The number of data for each hand hardnesslevel and microstructure are shown.

than 20 data there is an approximate probability of 0.95 of the population mean falling

within two standard errors of the sample mean, so the means of hand hardness for two

different microstructure classes differ at the 0.05 significance level when their whiskers do

not overlap.

The usefulness of hand hardness as an index of shear strength can be assessed from

the trends evident in Figure 6.4. For each of the four classes of microstructure, mean

strength increases significantly with hand hardness. Also, for hand hardness levels of F and

4F, the mean strength for decomposed and fragmented grains is significantly higher than

for precipitation particles. For each of the four hand hardness levels, the mean strength for

rounded grains is significantly stronger than for decomposed and fragmented grains. For

hand hardness levels of F, 4F and P, the mean strength of faceted crystals is less than for

rounded grains, and for 1F hardness, the reduction in mean strength for faceted crystals is

not significant. Consequently, as an index of shear strength, hand hardness is best

interpreted together with microstructure.

Since the area of the objects being pushed into the snow does not decrease

proportionally from fist to knife, hand hardness is an ordinal and not an interval measure.

The CAA's geometric scale for hand hardness (NRCC/CAA, 1989; CAA, 1995) is

intended to provide a graphical indication of resistance that better reflects quantitative

—but more time consuming—measures of the resistance such as is obtained with the ram

penetrometer (e.g. Martinelli, 1971). For this geometric scale, each major level of hand

hardness is plotted at twice the hardness-value of the preceding major level. Hence,

1-finger (1F) is considered to be twice as hard as four-finger (4F) and four times as hard

as fist (F), which is given an arbitrary value that allows all values to be plotted on a

particular graph. Using this doubling scale and the intermediate hardness levels such as

4F+, the Daniels strength is plotted against hand hardness in Figure 6.5 for the two classes

of microstructure with the most data. An approximately linear relationship is apparent

between Daniels strength and scaled hand hardness, which supports the use of the

doubling scale.

115

6.4 Characteristics of Persistent Slab Avalanches

6.4.1 Comparison of Reported and Investigated Dry Slab Avalanches

Before assessing the results of shear frame tests done near recent dry slab

avalanches, it is helpful to consider whether the avalanches selected for such tests are a

representative sample of the dry slab avalanches of concern to backcountry avalanche

forecasting. In addition to on-site investigations that included rutschblock tests and shear

frame tests on the failure plane as well as measurements of slab thickness, slope angle, etc.

(Appendix C), basic observations of many avalanches are available from daily occurrence

reports from ski guides. These occurrence reports include estimates of slab thickness,

avalanche width, slope angle, elevation, etc. and are based on a Canadian specification for

reporting avalanche occurrences (CAA, 1995). In this section, the slab thickness, slab

width and start zone inclinations from occurrence reports and on-site investigations are

compared (Table 6.2). The occurrence reports are limited to class 2 and larger avalanches

which are, by definition, large enough to injure, bury or kill a person. Smaller avalanches

are not reported consistently or completely and are of lesser importance to backcountry

forecasting and the present study. The avalanche occurrence reports used for Table 6.2 are

Figure 6.5 Shear strength plotted against scaled hand hardness showing approximatelylinear relationship for 126 layers of decomposed and fragmented particles and for 140layers of faceted crystals.

116

from the winters of 1990 to 1995 in the Cariboo and Monashee Ranges of the Columbia

Mountains.

Mean slab thickness and start zone inclinations from occurrence reports and

investigations are compared using a two-tailed t-test for unequal sample sizes and unequal

variances (Eq. 4.3). As shown in Table 6.2, there is no significant difference (p > 0.05)

between the mean thickness or width of the class 2 and larger slab avalanches reported by

ski guides and those investigated by researchers. However, the reported natural avalanches

started in significantly steeper terrain than the investigated natural avalanches. This

difference is not surprising since some natural avalanches start in very steep terrain which

can be difficult or unsafe to access for investigations. Also, the investigated skier-triggered

avalanches are significantly steeper than the reported skier-triggered avalanches. However,

since the reported start zone inclinations are estimated and the investigated start zone

inclinations are measured, the difference may simply be the result of inclinations being

under-estimated. During several investigations of reported avalanches, a tendency towards

under-estimated slope inclinations was noted.

117

Table 6.2 Comparison of Avalanche Characteristics from Occurrence Reports andOn-Site Investigations

CharacteristicReported Dry Slab

Avalanches1, Class 2 and Larger

Investigated Dry SlabAvalanches2

t-test

N Range Mean±S.D. N Range Mean±S.D. t p

Natural Slab Thickness (m) 286 0.1-2.0 0.53±0.32 14 0.4-1.5 0.59±0.29 -0.77 0.45

Slab Width (m) 239 8-1500 121±182 14 20-350 114±90 0.26 0.80

Start Zone Incline (o) 258 25-60 40±6.0 17 30-45 37±3.7 3.09 <10-2

Skier-Triggered

Slab Thickness (m) 36 0.2-1.2 0.55±0.24 51 0.1-1.5 0.46±0.26 1.66 0.10

Slab Width (m) 28 8-400 69±78 39 2-400 59±94 0.47 0.64

Start Zone Incline (o) 33 20-45 35±5.0 51 28-48 39±4.7 -3.67 <10-3

1 Reports are from Cariboo and Monashee Mountains near Blue River, BC.2 Investigations are from Cariboo, Monashee, Purcell and Selkirk Mountains.

6.4.2 Characteristics of Investigated Slab Avalanches

Aside from the differences in start zone inclination, the investigated slab avalanches

appear to be representative of the natural and skier-triggered slab avalanches in the

Columbia Mountains. Snowpack measurements and observations from the investigations

are summarized in Table 6.3 for natural and skier-triggered slab avalanches. Also,

Figure 6.6 shows the thicknesses of the slab, superstratum, weak layer and substratum for

the natural and skier-triggered dry slab avalanches in this study. An inclination of 38o is

typical of start zones for natural and skier-triggered dry slab avalanches in this study

118

Table 6.3 Characteristics of Investigated Dry Slab Avalanches from ColumbiaMountains 1990-95

Natural Skier-Triggered

N Range Mean ±S.D.

N Range Mean ±S.D.

Slab

Thickness (m) 14 0.37-1.5 0.59 ± 0.29 51 0.1-1.5 0.46 ± 0.26

Density (kg/m3) 10 101-259 189 ± 53 48 76-374 162 ± 70

Superstratum

Thickness (m) 14 0.03-0.30 0.15 ± 0.08 45 0.01-0.35 0.11 ± 0.08

Density (kg/m3) 3 70-232 143 ± 82 14 98-300 161 ± 54

Weak Layer

Thickness (mm) 41 20-30 23 ± 5 232 5-80 21 ± 16

Density (kg/m3) 2 180-230 205 ± 35 7 55-200 121 ± 49

Daniels Strength (kPa) 10 0.27-2.27 1.06 ± 0.60 48 0.04-2.123 0.57 ± 0.45

Substratum

Thickness (m) 12 0.01-0.34 0.13 ± 0.12 41 0.01-0.65 0.19 ± 0.14

Density (kg/m3) 1 294 294 12 106-450 212 ± 881 An additional 10 weaknesses were recorded as interfaces (thickness 0 mm)2 An additional 21 weaknesses were recorded as interfaces (thickness 0 mm)3 Includes high strengths from some remotely triggered avalanche discussed inChapter 8.

(Table 6.2) and in others (Perla,

1977; Williams and Armstrong,

1984, p. 201; Föhn, 1987a).

The dominant

microstructures for the

superstratum, weak layer and

substratum are summarized in

Figure 6.7. Although

approximately 15% of the

superstrata consisted of faceted

crystals, microstructures such as

rounded grains, crusts and

decomposed and fragmented

precipitation particles were more

common. Depth hoar and surface

hoar crystals were not observed

in the superstrata. Faceted crystals

119

Figure 6.6 Cross section of typical dry slab avalanches. Layer thicknesses are measuredvertically.

Figure 6.7 Relative frequency of microstructures forsuperstratum, weak layer and substratum of dry slabavalanches in Columbia Mountains, 1990-95.

and surface hoar were commonly observed in the weak layers but crusts were never

reported as a weak layer. Faceted crystals were more common in weak layers than depth

hoar, indicating that the earlier products of kinetic metamorphism are often critically

weak, a finding consistent with Bradley and others (1977a, b) and Adams and Brown

(1982). The weak layers are more likely to consist of faceted crystals or surface hoar than

are the superstrata, and the substrata consist of a wide assortment of microstructures.

Crusts occurred more often in the substrata than in the superstrata.

The microstructures of the superstrata, weak layers and substrata are shown

separately for natural and skier-triggered dry slab avalanches in Figure 6.7. Less

metamorphosed forms such as precipitation particles and decomposed and fragmented

precipitation particles were observed more often in the superstrata of skier-triggered

avalanches than in natural avalanches. The substrata for skier-triggered slabs also show the

same bias toward less metamorphosed microstructures compared to natural slab

avalanches. This is not surprising since skiers, and guides in particular, ski test slopes with

shallower, younger slabs, and avoid slopes with deeper, older slabs that might be unstable

and consequently, more dangerous. For natural avalanches, faceted crystals were found in

the weak layers more often than surface hoar, whereas the opposite was true for

skier-triggered avalanches. This association of surface hoar with skier triggering is

consistent with a study of fatal accidents in Canada (Jamieson and Johnston, 1992a) in

which 41% of the identified weak layers consisted of surface hoar.

The distribution of hand hardness for the superstratum, weak layer and substratum is

shown in Figure 6.8 for natural and skier-triggered dry slab avalanches in the Columbia

Mountains. Although the hand hardness levels are scaled (NRCC/CAA, 1989; CAA,

1995), the distributions are summarized with the median, and 10, 25, 75 and 90th

percentiles which are suited to ordinal data. For natural avalanches and skier-triggered

avalanches, the mean hardness of the weak layers (4F) are less than for the superstrata

(1F) or substrata (P- for natural avalanches and 1F for skier-triggered avalanches). As

well, based on the CAA’s doubling scale, the hardnesses of the superstrata and the

substrata are more variable than the hardness of the weak layers. Although a softer weak

120

layer sandwiched between two harder layers is common for slab avalanches, similar

“sandwiches” also occur in stable snowpacks. Nevertheless, recognition of such hardness

sandwiches is helpful for stability evaluation (e.g. Fredston and Fesler, 1994, p. 56-57).

Subsequent sections of this chapter assess shear frame stability indices for natural

avalanches (Section 6.5 and 6.6) and for skier-triggered avalanches (6.7, 6.8 and 6.9).

6.5 Predicting Natural Avalanches on Test Slopes

In this section, a stability index for natural avalanches is compared with natural

avalanche activity on avalanche slopes tested with the shear frame. Although it is

impractical for avalanche safety operations that forecast for large backcountry areas to test

weak layers in numerous starting zones with the shear frame, a comparison of avalanche

activity for low and high values of the stability index is a useful way of assessing whether

the index can discriminate between stable and unstable slopes (Föhn, 1987a).

The slope-specific stability index for natural avalanching is

121

Figure 6.8 Resistance for superstratum, weak layer and substratum of dry slabavalanches in Columbia Mountains, 1990-95.

(6.3)SN =Σφ

ρgh sin Ψ cos Ψ

where Sf is the Daniels strength adjusted for the effect of the normal load due to the

overburden, r is the average slab density, h is the vertical thickness of the slab and Y is the

slope inclination in the start zone. This index differs from S developed by Föhn (1987a)

only in that the effect of normal load on shear strength varies with the microstructure of

the weak layer, based on data presented in Section 4.11. Since the effect of normal load on

persistent microstructures is taken to be negligible based on field data presented in Figure

4.7, SN is lower and hence potentially more conservative than S for persistent slabs,

especially for deep slabs which apply high normal loads to the underlying weak layer. (An

error analysis for this index is presented in Appendix B.)

Shear frame tests were done within 2 days of 10 natural dry avalanches involving

persistent slabs and one involving a non-persistent slab. The mean slab thickness of these

avalanches is plotted against the stability index SN in Figure 6.9. In each case, SN is based

122

Figure 6.9 Values of SN for slopes that did and did not avalanche naturally.

on shear frame tests of the failure plane. For comparison, Figure 6.9 also includes points

from 29 persistent slabs and 33 non-persistent slabs that did not avalanche, in which case

the weak layer judged most likely to fail was tested with the shear frame.

Since there is only one non-persistent slab that failed naturally, SN's ability to predict

the natural stability of non-persistent slabs cannot be determined.

However, 10 of the 39 persistent slabs failed naturally. Values of SN for the 10

persistent slabs that released naturally range widely from 0.9 to 8.5. The four natural

avalanches for which SN > 4 are labelled with the location in Figure 6.9. It appears that

increased air temperatures contributed to the slab failures or that subsequent cooling

contributed to the surprisingly high values of SN. At the Roller Coaster avalanche on

1 February 1993, a crust had developed on the surface of the slab and bed surface,

indicating that at least one melt-freeze cycle had occurred after the avalanche and before

the shear frame tests two days later, rendering the shear frame results questionable. The

Arrowhead avalanche occurred at 2700 m elevation during marked warming and snowfall

on 2 March 1994 when freezing levels and rain rose to approximately 2000 m. The tests

were done a day later when the air temperature in the start zone had dropped to -6.7oC.

The Elk avalanche occurred at 2500 m elevation on 4 March 1994 during an unusually

warm period when the freezing level rose to 2000 m. The shear frame tests were done a

day later when the air temperature had dropped to -13.5oC. Similarly the Sherbrooke

avalanche occurred on 28 December 1994 when air temperatures rose to -2oC and the

shear frame tests were done 2 days later when the air temperature had dropped to -21oC.

The Elk and Sherbrooke avalanches occurred during warming without substantial loading

and the Arrowhead avalanche occurred during warming and loading by snowfall.

Nevertheless, SN does not appear promising for predicting natural avalanches in which

warming contributes to slab failure by increasing the shear strain rate without increasing

the shear stress. This is not surprising since the stability index SN is a critical stress failure

criterion (Eq. 6.3) whereas laboratory studies of snow in tension (Salm, 1971; Narita,

1980, 1983) show that the peak stress for ductile failure depends on the strain rate and

Brown and others (1973) showed that the peak stress depends on the strain history.

123

The merit of SN for predicting natural avalanches due to loading by precipitation

could also be questioned because the critical stress for ductile failure depends strongly on

the strain rate which is not part of the shear frame stability index2. Yet, shear frame

stability indices have already proven themselves (Schleiss and Schleiss, 1970; Föhn,

1987a; Jamieson and Johnston, 1993a) to be useful for predicting natural avalanches, most

of which are caused by loading due to precipitation.

However, the critical values of shear frame stability indices which are based on

brittle strength of small areas are questionable since natural avalanching involves initial

ductile failure of large areas. Schleiss and Schleiss (1970) found Σ/ρgh to be critical at

approximately 1.5. Jamieson and Johnston (1993a) found a similar index for natural

avalanching to be critical between 2.8 and 3. Further, in Föhn’s (1987a) study in the Swiss

Alps, values of SN for 18 dry slabs that failed averaged 2.3 and SN was less than 1.5 for

only three of these slabs, suggesting a critical value greater than 2.3. In Figure 6.9, only 10

values of SN were obtained for persistent slabs that failed and four of these can be rejected

based on temperature effects. However, a critical value near 2 appears likely since SN < 2

for five of the six natural avalanches not attributed to warming and SN > 2 for 24 of the 29

slabs that did not avalanche naturally.

The number of slabs that avalanched naturally and were tested with shear frames is

very limited partly because access to fresh dry slab avalanches within a day or two of slab

failure is restricted by safety considerations, helicopter logistics, poor flying weather and

snowmobiling difficulties. However, the primary reason for this limitation is the fact that

research technicians were directed to try to access skier-triggered slab avalanches in

2 A critical strain rate failure criterion might be more promising since, based on the

work of McClung (1977) and Narita (1980), Bader and Salm (1990) showed that the

critical strain rate for the low density snow associated with weak layers was approximately

10-4 s-1. However, Sommerfeld's (1979) strain gauge on an avalanche slope was removed

and presumably damaged by the first avalanche. A practical strain gauge for avalanche

slopes and stability index based on strain rate have not yet been developed.

124

preference to natural avalanches whenever there was a choice because skier-triggered

avalanches cause many more backcountry fatalities than do natural avalanches.

The next section assesses a variation of SN using reported natural avalanche

occurrences from surrounding slopes where more data were available than for

investigated avalanche slopes tested with the shear frame.

6.6 Predicting Natural Avalanches of Persistent Slabs on SurroundingSlopes

In this section, natural avalanche activity over a large area is compared with a

stability index measured at a safe study site. Although most of the dry slab avalanches

were within 10-15 km from the study site, some were 30 km away. The study sites ranged

in inclination from 0 to 32o, were rarely exposed to avalanches from slopes above, and

were generally sheltered from the wind. However, aside from wind effects, sites were

selected that had a snowpack similar to surrounding avalanche terrain. The study sites

were near treeline and ranged in elevation from 1900 to 2300 m. Most avalanche start

zones in the study areas are between 1700 and 3000 m. Most snow storms are

accompanied by wind from the south, south-west or west which increases the amount of

snow deposited on the north, north-east and east slopes. The inclined study sites face

north, north-east and east as do many of the slide paths on which avalanches were

reported.

Since the objective of this section is to apply a stability index to slopes of various

inclinations many km away from the study slope, a slope angle of Ψ = 38o that is typical of

slab avalanche start zones (Table 6.2) is used in Eq. 6.3 to obtain SN38.

Between the low values of a stability index that are associated with unstable slabs

and the high values that are associated with stable slabs, there is a critical or transitional

value. In practice, the transition consists of a band of values because of differences in

snow conditions between avalanche starting zones and study sites and because of the

variability of shear strength and overburden measurements. Since stability indices are used

in conjunction with other observations, the width of the transition band is based on the

125

90% confidence band for the stability indices. The width of this band can be approximated

from the shear strength measurements which are more variable than the density

measurements. Typically, values of SN38 are based on 12 shear frame tests in a study site

that have an average coefficient of variation of 15% (Section 4.3). This corresponds to a

90% confidence band approximated by ±10% of the transitional value.

Persistent weak layers underlying thick slabs generally change in strength more

slowly than persistent weak layers underlying shallow slabs. For this reason, persistent

weak layers were tested with the shear frame approximately twice per week when the slab

depths in the study sites were less than approximately 0.6 m and once per week when the

persistent weak layers were more deeply buried.

Dry slab avalanche activity can be compared with values of the stability index on

those days that persistent weak layers were tested with the shear frame. However,

avalanche activity can also be compared with stability indices between test days by linearly

interpolating the strength Σ∞ between test days and estimating the load σv = ρgh daily

based on precipitation recorded at the nearest weather station. For example, if the shear

strength and load were measured at a particular study site on the ith and kth days, then the

load on the jth day is

(6.4)σv,j = σv,j−1 +HNWj

Σ ik HNW

(σv,k − σv,i)

where i < j < k and HNW is the water equivalent of the precipitation over the previous 24

hours. Values of (σv,k - σv,i)/ΣikHNW vary with the study site. At Mt. St. Anne in the

Cariboos, the study plot receives an average of 0.013 kPa of load in the study plot per

mm of water in the precipitation gauge located 100 m away. Similarly, Sam’s study site in

the Monashee Mountains receives an average of 0.015 kPa of load per mm of water in the

Mt. St. Anne precipitation gauge 23 km away. In the Purcell Mountains where the height

of snowfall within the previous 24 hours, HN, was used in lieu of HNW which requires a

melting precipitation gauge or regular weighing of a column of snow from the previous 24

hours, 0.012, 0.016, 0.021 and 0.032 kPa of load are typically received at the Vermont

(1600 m), Elk (2250 m), Pygmy (2150 m) and Rocky (1900 m) study sites respectively for

126

each cm of snowfall recorded at the Bobby Burns Lodge (1370 m). The Rocky study site

is located in the Crystalline Valley which is known for heavier snowfall than nearby

valleys.

While precipitation and hence increase in load could be estimated daily for particular

study sites based on precipitation at the nearest weather station, the changes in strength

cannot be estimated based on meteorological parameters or easily measured snowpack

properties. This means that at present the only way to determine a shear frame stability

index is to access a study site, dig a pit and perform the shear frame tests, a practice that

would be impractical for most backcountry avalanche forecasting operations to do on a

daily basis. However, by interpolating load between test days, the merit of extrapolating

stability indices to surrounding terrain can be assessed.

Extrapolating to surrounding terrain with a stability index is not new. The highway

avalanche safety programs for the Trans-Canada Highway through Rogers Pass (Schleiss

and Schleiss, 1970) and for BC Highway 3 through Kootenay Pass use the ratio Σ100/ρgh

for surrounding terrain. However, neither program has been able to verify the merit of this

ratio mainly because operational staff only perform the shear frame tests when other

forecasting factors indicate low stability (Salway, 1976). Jamieson and Johnston (1993a)

obtained shear frame data from Mt. St. Anne during days in which dry slab avalanches

were reported and days in which no such avalanches were reported. The shear frame tests

were done on the weak layer that appeared to be least stable at the study site. They used

the ratio Σ250/ρgh as well as the index S35 which uses the normal load correction for

rounded grains (Roch, 1966b; Föhn, 1987a). Using empirically chosen critical levels and a

±10% band for transitional stability, the predictive merit of the indices was indicated by

the fact that the values were below the transition band on approximately 84% of days in

which natural dry slab avalanches were reported and above the transition band for 73% of

the days in which no natural dry slab avalanches were reported. However, that study

correlated S35 for the weak layer that appeared least stable in the study plot with all dry

natural slab avalanches that were reported to occur on the day of the shear frame tests. S35

127

was not interpolated between test days, and neither the tested weak layers nor the failure

planes for the natural dry slab avalanches were restricted to persistent weak layers.

In general, selecting a critical level that is too high will result in excessive false

predictions of instability, and a critical level that is too low will result in excessive false

predictions of stability. The latter type of prediction errors are called false stable

predictions and are potentially far more serious because they can contribute to an unstable

slope being judged stable. In the Swiss Alps, values of SN for 18 dry slabs that avalanched

naturally averaged 2.3 (Föhn 1987a), suggesting a critical value greater than 2.3.

However, when extrapolating to surrounding terrain a higher critical value may be

appropriate because of the greater variability in snowpack properties over larger areas and

the need to reduce the proportion of false stable prediction errors. Jamieson and Johnston

(1993a) empirically determined critical values of S35 that averaged 3.0 for the level study

plot and 2.8 for a nearby study slope at Mt. St. Anne. In the following studies of SN38, a

critical value of 2.8 is used to assess the merit of SN38 for discriminating between

avalanche and non-avalanche days.

Since SN38 is a ratio of shear strength to shear stress, increases in SN38 over time

indicate that the shear strength of the weak layer is increasing faster than shear stress due

to slab weight, whereas decreases in SN38 indicate that the shear stress due to slab weight

is increasing faster than the shear strength of the weak layer. Decreases in shear strength

due to metamorphic weakening of the weak layer are not common in the Columbia

Mountains and measurable decreases in load due to sublimation exceeding precipitation

are also not common; therefore most decreases in SN38 are due to increases in slab weight

and shear stress caused by precipitation.

Due to varying amounts of snowfall (load) between study sites and start zones,

extrapolation of stability from a study site to a start zone might not seem promising.

However, snow layers under heavier slabs tend to densify and strengthen faster than an

initially similar layer under a lighter slab (Armstrong, 1980). This tendency reduces the

differences between stability indices for particular weak layers from areas with different

loads.

128

When attempting to correlate a stability index of a particular weak layer in a study

site with avalanche activity over a wide area, there may be uncertainty associated with the

identification of the failure plane of the avalanches. Some fresh avalanches cannot be

approached safely. More often, operational staff do not have time to verify the failure

plane by visiting the crown or flanks. Frequently, an observer identifies the failure plane

from a distance of up to several hundred metres away based on the depth of the crown and

flank fractures and on knowledge of weak layer depths from snow profiles often within a

few km of the avalanche.

Typically, two to four persistent weak layers are repeatedly tested with the shear

frame over a period of two or more weeks each winter at each of the two main forecast

areas, i.e. the Purcell Mountains near Bobby Burns Lodge and the Cariboo and Monashee

Mountains near Blue River, BC. However, not all of these persistent weak layers produce

sufficient avalanches for comparison with stability indices such as SN38. From those

persistent layers that were monitored, there were three from the Purcell Mountains, three

from the Cariboo and Monashee Mountains, and two from Jasper National Park that

produced natural dry slab avalanches on at least two days. Based on testing of these eight

layers, the ability of SN38 to discriminate between avalanche and non-avalanche days is

assessed in the following three subsections.

6.6.1 Purcells

During the winter 1992-93, shear frame tests were done at two level study plots:

Rocky at 1900 m and Pygmy at 2150 m, and at 25-35o slopes adjacent to these plots. At

each of these sites a layer of surface hoar buried on 18 January 1994 was tested. Intervals

between testing ranged up to 14 days, partly since a helicopter was unavailable for site

access during two 1-week periods. The only natural dry slab avalanches that were

reported to have started on the surface hoar layer occurred on 30 January 1993 and 4

February 1993 (Figure 6.10). Although shear frames tests had not been started yet on the

surface hoar at Rocky Plot, the index SN38 was below 2.8 at Pygmy Plot, Pygmy Slope and

Rocky Slope on 30 January 1993 when approximately five dry slab avalanches started on

the surface hoar layer. However, by 4 February 1993 when the last avalanche started on

129

the surface hoar

layer, SN38 was below

2.8 only at the Rocky

Slope. For three to

four weeks, there

were no natural dry

slab avalanches

reported and SN38

remained above 2.8.

During 6 March to

15 March, natural

dry slab avalanches

occurred when SN38

was below 2.8 at

Rocky Plot, Rocky

Slope and Pygmy Plot.

Further, during 12

February to 14 February when SN38 at Pygmy Plot, Pygmy Slope and Rocky Slope was

above 2.8, no dry slab avalanches occurred. Although a few of these avalanches between 6

March and 15 March started on the surface hoar layer buried 10 February 1993, the failure

plane was not reported for most of these avalanches. Note that the afternoon air

temperature was above freezing at 1370 m when the avalanches occurred on 6 March and

7 March 1993. Nevertheless, based on Figure 6.10, a critical value of 2.8 for SN38 appears

useful for discriminating between stable and unstable periods, even though it failed to

predict the single avalanche on 4 February 1993.

A second layer of surface hoar was buried at all four study sites on 10 February

1993. However, cold weather between 14 February and 26 February (Figure 6.11) caused

a strong temperature gradient within the snow on top of the surface hoar, resulting in

cohesionless faceted snow instead of a cohesive slab. There were numerous small loose

130

Figure 6.10 Stability trend for natural avalanches on surfacehoar buried 19 January 1993. in the Purcell Mountains.

avalanches but no slab

avalanches during this

period. The marked

decrease in SN38 at all

sites on 5 March 1993

was due to heavy

snowfall. During 6

March and 11 March,

SN38 was below 2.8 at all

sites and dry slab

avalanches occurred.

Three of these were

reported to have failed

on the surface hoar layer.

No dry slab avalanches

occurred between 12 March

and 15 March while SN38 increased at Pygmy Plot, Pygmy Slope and Rocky Plot and was

above 2.8 at Rocky Plot. On 16 March, more than 10 dry slab avalanches started on the

surface hoar and SN38 dropped below 2.8 at the three sites at which the surface hoar was

still being tested as a result of renewed snowfall. As with the previous figure, a critical

value of 2.8 for SN38 appears useful for discriminating between stable and unstable periods

in Figure 6.11.

In subsequent seasons, persistent weak layers were tested with the shear frame at

Pygmy Slope and Rocky Slope. However, the testing in the level plots was discontinued

because the study slopes appeared to work as well as the level plots, and because any

effect of creep within the snowpack on the strength of surface hoar would better reflect

the strength changes of surface hoar in start zones.

In the winter of 1993-94, a layer of surface hoar and/or faceted crystals was buried

on 6 February in the Purcell Mountains. A new study plot (Elk) representative of

131

Figure 6.11 Natural stability trend for surface hoar layerburied 10 February 1993 in the Purcell Mountains.

shallower snowpack

areas more likely to

produce and retain

persistent weak layers

was added. The surface

hoar layer was monitored

with the shear frame at

the Pygmy, Rocky and

Elk study sites where it

initially consisted of

1-2 mm faceted crystals,

2-3 mm surface hoar and

6-10 mm surface hoar,

respectively. Throughout

the testing, SN38 at the Elk

Plot remained above

values from the Rocky and Pygmy sites (Figure 6.12). Natural dry slab avalanches were

reported to have started on the surface hoar on 14 February 1994, 3 March, 4 March and

6 March. SN38 was below 2.8 at the Elk Plot when the avalanches occurred on 14 February

and the air temperature at Bobby Burns Lodge was above 0°C on 3 March and 4 March

when the other avalanches occurred. However, there were many days in which SN38 from

the Elk Plot was below 2.8 and no natural dry slab avalanches on the surface hoar were

reported. Throughout the test period at the Pygmy site and for most of the period at the

Rocky site, SN38 was below 2.8. While SN38 from these two sites was less than 2.8 on the

days that the natural dry slab avalanches were reported to have started on the surface

hoar, there were many days in which SN38 was below 2.8 and no natural dry slab

avalanches were reported to have started on the surface hoar layer, so a critical value of

SN38 = 2.8 was not very effective at discriminating between stable and unstable periods.

132

Figure 6.12 Stability trend for a layer of surface hoar buried 6February 1994 in the Purcell Mountains.

6.6.2 Cariboos and Monashees near Blue River

On 29 December 1993, a layer of surface hoar was buried in the Cariboo and

Monashee Mountains near Blue River, BC. However, size of the surface hoar varied

considerably throughout the forecast area. At Sam’s Study Plot, the surface hoar was

2 mm in size and at the Mt. St. Anne Study Plot the surface hoar was 6-9 mm in size,

larger than in surrounding terrain. Not surprisingly, SN38 for this layer was lower at the Mt.

St. Anne Study Plot than at Sam’s Study Plot throughout the following 5 weeks as shown

in Figure 6.13. Since these two study plots receive similar amounts of precipitation, the

difference in SN38 values indicates that the 2 mm surface hoar at Sam’s Plot was

substantially stronger than the 6-9 mm surface hoar at Mt. St. Anne throughout the

period. During the

period 1 January

1994 to 4 January

1994, natural dry

slab avalanches

started on the surface

hoar in the Cariboo

Mountains when SN38

at both sites was at

its lowest values.

However, after 4

January SN38

remained below 2.8

at Mt. St. Anne and

none of the natural

dry slab avalanches

were reported to

have started on the

surface hoar. The size of

133

Figure 6.13 Stability trend for a layer of surface hoar buried29 December 1993 in the Cariboo and Monashee Mountainsnear Blue River, BC.

the surface hoar when it was

buried appears to be critical.

In the Mt. St. Anne Study

Plot where it was larger than

in surrounding terrain, SN38

remained low and failed to

predict the absence of

avalanches on the surface

hoar. In Sam’s Plot, where the

surface hoar was relatively

small, SN38 was above 2.8

from 1 January to 3 January

when natural dry slab

avalanches were starting on

the surface hoar layer.

On 5 February 1994, a layer

of surface hoar was buried in the

Cariboo and Monashee Mountains.

Although most of the natural dry slab avalanches that started on the surface hoar occurred

when SN38 at Mt. St. Anne and Sam’s Plots was minimal (Figure 6.14), the absence of

natural dry slab avalanches starting on the surface hoar after 17 February was not

predicted since SN38 remained below 2.8 throughout the period at both sites.

A layer of large surface hoar was buried in the Cariboo and Monashee Mountains

near Blue River on 7 January 1995. Only 3 natural dry slab avalanches were reported on

this surface hoar layer and these occurred between 17 January and 19 January 1995 as

shown in Figure 6.15. Although the index SN38 reached 3.1 on 18 January, it was below

2.8 on the preceding and following days when the other avalanches started on the surface

hoar layer. No other natural dry slab avalanches were reported to have started on the

surface hoar although SN38 dropped below 2.8 from 6 February to 12 February. During the

134

Figure 6.14 Stability trend for a layer of surfacehoar buried 5 February 1994 in the Cariboo andMonashee Mountains near Blue River, BC.

period from 10 January to 18

February, SN38 was generally

effective at discriminating

between periods with and

without natural dry slab

avalanches that started on the

surface hoar layer.

6.6.3 Rocky Mountains

A layer of snow that fell

on a crust in Jasper National

Park was weakened by faceting

during cold weather during

November and December 1993.

This resulted in a layer of facets

approximately 0.1 m thick

located approximately 0.1 m above the ground. On 6 days between 1 December 1993 and

9 January 1994 this layer was tested with a shear frame at the Sunwapta Study Plot

(2000 m). Based on these tests, SN38 ranged between 0.9 and 1.4 staying well below 2.8

(Figure 6.16). Avalanche observations were conducted along the Icefield Parkway.

Natural dry slab avalanches that started in these facets were reported intermittently

between 4 December 1993 and 9 January 1994 and continued after the shear frame tests

on this weak layer were discontinued. Since there were many days in which SN38 < 2.8 and

no avalanches occurred, SN38 did not effectively predict days without avalanches.

However, the snow in the start zones in this area of Jasper National Park is strongly

affected by wind whereas the Sunwapta Study Plot is sheltered from wind. Possible

explanations include:

1. A stability index like SN38 based on shear frame tests in a sheltered study plot is not

relevant to surrounding wind-affected start zones.

135

Figure 6.15 Stability trend for a layer of surface hoarburied 7 January 1995 in the Cariboo and MonasheeMountains near Blue River, BC.

2. SN38 < 2.8 represents a condition that is necessary for natural avalanches but not

sufficient.

3. The stability index, SN38, that is based on the ratio of shear strength to shear stress is

better suited to the failure of thin weak layers more common in the Columbia

Mountains than to the failure of thick weak layers common in the Rocky Mountains.

On 8 February 1994, a layer of surface hoar was buried in Jasper National Park (and

2-4 days earlier in the Columbia Mountains to the west). Shear frame tests were

conducted on this layer in the Sunwapta Study Plot on 5 days between 10 February and 14

March 1994. During the period 10 February to 12 February, SN38 was above 2.8 and no

natural dry slab avalanches were reported to have started on the surface hoar layer

(Figure 6.17). Except for 27 February, SN38 remained below 2.8 until tests were

discontinued on 15 March 1994. Natural dry slab avalanches started on the surface hoar

136

Figure 6.16 Stability trend for a layer of facets formed inOctober 1993 in Jasper National Park.

during 9 of the 30 days between 13 February and 15 March that SN38 was below 2.8. As

with other persistent weak layers from Jasper National Park, all the avalanches occurred

when SN38 was below 2.8, but there were 21 days without avalanches during which

SN38 < 2.8.

6.6.4 Summary for Natural Stability Indices

When considering the effectiveness of a stability index for discriminating between

days with and days without natural dry slab avalanches, it must be noted that no single

variable provides a sound basis for predicting natural dry slab avalanches. The index SN38

and the critical value 2.8 appeared to have some predictive value for the stability of the

persistent weak layers buried on 18 January 1993, 10 February 1993 and 6 February 1994

in the Purcell Mountains. However, SN38 was not effective for the surface hoar layers

137

Figure 6.17 Stability trend for a layer of surface hoar buried 8 February 1994in Jasper National Park.

buried on 29 December 1993 and 5 February 1994 in the Cariboo and Monashee

Mountains near Blue River, BC. SN38 appeared to be a useful predictor of the natural dry

slab avalanche activity for the surface hoar layer buried on 7 January 1995 in the Cariboos

and Monashees near Blue River. The effectiveness of SN38 for the Rocky Mountains

remains unclear. The studies at Jasper National Park are confounded by strong wind

effects in start zones.

Although Schleiss and Schleiss (1970) found the ratio Σ/ρgh useful for weak layers

under the snow from the most recent storm, and Jamieson and Johnston (1993a) found S35

useful for natural dry slab avalanches, many of which started in weak layers buried only by

the most recent storm, SN38 which is similar to Σ/ρgh and S35 appears to have limited

predictive value for persistent layers that remain weak for several storms. The difference

probably lies in the depth and persistence of the weak layers. Persistent layers such as

surface hoar and facets are the failure planes for many thick natural avalanches, whereas

weak layers of precipitation particles are usually the failure plane for shallower slab

avalanches mainly involving snow from the most recent storm. This supports the unproven

statement by Schleiss and Schleiss (1970) that the shear frame test is most effective for

weak layers of “new snow” (precipitation particles).

However, few fatal avalanches start naturally; most are triggered by people

(Seligman, 1936, p.336; Jamieson and Johnston, 1993b). The next three sections assess

stability indices for skier triggering.

6.7 Predicting Skier-Triggered Avalanches on Test Slopes

In this section, a stability index for skier-triggering is assessed using the results of

skier-testing on avalanche slopes. As in section 6.5, the failure plane or potential failure

plane is tested with the shear frame at a site judged typical of the start zone.

As noted in Chapter 2, Föhn (1987a) derived a stress term for the stress induced in a

weak layer by a static skier, ∆σxz, and included it in the denominator of the stability index

138

S' (Equation 2.7). Replacing the normal load adjustment in the numerator of S' by

φ(Σ∞,σzz) which depends on microstructure (Section 4.11) yields

(6.5). SS =Σ∞ + σzzφ(Σ∞, σzz)

σxz + ∆σxz

Slopes were skier-tested before the shear frame tests were performed. To reduce the

hazard to the tester during unstable conditions and especially when the slab thickness

exceeded 0.3 m, short slopes were selected. Occasionally, these short slopes “failed” but

did not release an avalanche, as indicated by a tension crack through the slab near the top

of the slope and a flank crack down along one or both sides of the slab.

In a few cases in which the slab did not fail there were two weak layers. When time

permitted, both layers were tested with the shear frame resulting in two data points for the

same slope on the same day.

The slab thickness in the start zone hSZ is plotted against SS for 63 persistent slabs

(square symbols) and 26 non-persistent slabs (circular symbols) in Figure 6.18. Slabs that

failed are marked with filled symbols, and slabs that did not fail are marked with unfilled

139

Figure 6.18 Stability index SS for skier-tested avalanche slopes.Slabs triggered from more than 50 m away from the displaced slabare marked with square around the symbol.

symbols. Symbols indicating slabs that were remotely triggered from a distance of at least

50 m from the displaced slab are surrounded by a square.

Within the unstable range SS < 1, 81% (25/31) of the persistent slabs and 69%

(9/13) of the non-persistent slabs failed. Within the band of transitional stability used by

Föhn (1987a), 1 ≤ SS ≤ 1.5, 70% (7/10) of the persistent slabs and 13% (1/8) of the

non-persistent slabs failed. In the stable range, SS > 1.5, 27% (6/22) of the persistent slabs

and 0% (0/5) of the non-persistent slabs failed. For persistent weak layers, the proportion

of slabs that failed decreased from 81% to 70% to 27% for the stable, transitional and

unstable ranges respectively. For non-persistent weak layers, the proportion of slabs that

failed decreased from 69% to 13% to 0% for the stable, transitional and unstable ranges

respectively. These decreases in the proportion of slab failures for increasing values of a

stability index are measures of the effectiveness of the index.

Prediction errors are defined as false unstable if the stability index < 1 and the slabs

did not fail when skier-tested, or false stable if the stability index > 1.5 and the slab was

skier-triggered. In terms of prediction errors, there were 6/31 (19%) false unstable

predictions for persistent slabs and 4/13 (31%) for non-persistent slabs. Also, there were

6/22 (27%) false stable results for persistent weak layers and none (0/5) for non-persistent

layers. These proportions of predictions errors are similar to the 25% reported by Föhn

(1987a) for slabs with various triggers. This shows that Föhn’s (1987a) study is repeatable

and that skier-stability indices such as S' or SS are equally effective in the Columbia

Mountain snowpack as in the Swiss Alps despite differences in climate and snowpack.

While both false stable and false unstable results are prediction errors, it is more

important to minimize false stable results than false unstable results. If a stability index like

SS strongly influenced decisions about where and where not to ski, then false stable

predictions would have far greater consequences and costs (e.g. serious accidents, medical

costs, legal fees, etc.) than false unstable results which would only result in stable slopes

being avoided and, at worst, customer dissatisfaction. If the costs of a false stable

prediction are k times greater than those of a false unstable prediction, then a slope should

be avoided if the probability of a slab failure exceeds 1/(k + 1) (Blattenberger and Fowles,

140

1995a, b). Since k probably exceeds 100, slopes should be avoided that have a probability

of a hazardous avalanche as low as 10-2.

Although slope-specific shear frame stability indices based on shear frame tests are

probably too time-consuming for backcountry skiing operations that ski many slopes per

day, the success of SS (Figure 6.18) is compared with a refined stability index in the next

section which is the basis for extrapolated stability indices in Section 6.9 and Chapter 9.

Shear frame tests are, as a matter of practice, done where slab properties are judged

typical of the start zone. For avalanches that are triggered remotely, it is possible that the

snowpack properties at the trigger point differ substantially from those at the site of the

shear frame tests. In particular, three of the six false stable results involve remote

triggering as indicated by a square surrounding the symbol in Figure 6.18. Remote

triggering is discussed further in Chapter 8.

6.8 A Skier Stability Index for Soft Slabs

One of the assumptions inherent in SS is that the skier's weight is applied at the snow

surface. However, in the soft, low-density snowpack typical of the Columbia Mountains,

skis typically penetrate the snow surface by 0.3 m. For ski penetration during skiing, PK,

resulting in the skis being hSZ - PK above the weak layer, the stress induced by a static skier

(Eq. 2.8) on a slope of inclination ΨSZ is

(6.6)∆σ xz = 2L cos αmaxsin2αmaxsin(αmax + ΨSZ)π(hSZ − PK)cos ΨSZ

The effect of this adjustment for ski penetration on combined shear stress due to the

slab and skier, σXZ + ∆σ'XZ is shown in Figure 6.19 for penetrations of 0.0, 0.2 and 0.4 m.

Using a density at depth v (measured vertically) of 125 + 150v kg/m3 which is shown to be

typical of the Columbia Mountains in the next section, the shear stress in the weak layer

due to the slab of thickness h is

(6.7).σxz = g sin Ψ cos Ψ ∫0

h

(125 + 150v)dv

For weak layers within 0.7 m of the surface, such penetrations substantially increase

the total shear stress on the weak layer (Figure 6.19), indicating the relevance of

141

modifying SS for ski

penetration.

Skier triggering of a

slab is most likely when ski

penetration is greatest.

While skiing in powder

snow, ski penetration

reaches a maximum, PK,

when a skier pushes down

with the skis between

turns. Since PK cannot

readily be measured and

depends on the weight of

the skier, the area and

stiffness of the skis as well

as skiing technique, it was

decided to estimate the maximum penetration during skiing for an average skier.

Estimation based on the resistance profile of hand hardnesses was considered, but this

would have required additional measurements involving an ordinal measure of hardness,

and would probably have made the stability index calculations too complicated for many

hand-held calculators. For these reasons ski penetration is estimated based on slab density

which was available from measurements of slab weight per unit area (load) and slab

thickness which were already required for indices such as SN and SS.

6.8.1 Density Profiles

Snow density usually increases with depth, although some wind slabs are exceptions

to this generalization. Assuming a linear increase of density with depth, the density at

depth v is

(6.8)ρv = ρ0 +∆ρ∆v

v

142

Figure 6.19 Effect of ski penetration on skier-inducedstress.

where ρ0 is the density at the surface. This assumption of linearity was assessed using

density profiles from the Columbia Mountains. From 128 profiles at least 0.5 m deep that

were observed during the winters of 1993-95, there were 45 with mean slab density less

than 160 kg/m3 and 42 with mean slab density greater than 200 kg/m3. For each of these

groups of slabs, the densities are averaged at increments of 0.1 m between 0.1 m and

0.6 m as shown in Figure 6.20. Increases in density with depth appear linear and linear

regressions yield ∆ρ/∆v = 143 kg/m4 for the low density slabs and 167 kg/m4 for the high

density slabs. Since ∆ρ/∆v

shows little dependence on

mean slab density, a nominal

value of ∆ρ/∆v = 150 kg/m4 is

subsequently used for all slabs.

Since the mean slab

density, ρ, of an idealized slab

occurs at v = h/2

ρ = ρ0 +∆ρ∆v

(h/2)

(6.9)

Using ρ0 from

Equation 6.10 and

evaluating Equation 6.9 at

v = 0.3 m which is typical of skiing penetration, the density 0.3 m below the surface can be

estimated

(6.10)ρ30 = ρ +∆ρ∆v

(0.3 − h/2)

where the mean slab density, ρ, is obtained from σV/h where σV = ρgh is the load (slab

weight per unit area) measured with core samples or a density profile (Section 3.4).

143

Figure 6.20 Profiles of averaged densities for high and lowdensity slabs from the Columbia Mountains.

6.8.2 Estimating Ski Penetration from Average Slab Density

As part of the rutschblock study (Chapter 7), measurements of ski penetration were

obtained after gently stepping onto previously undisturbed snow, and after two jumps on

the same spot. The mean of these two measurements is taken as the skiing penetration, PK.

Measurements were taken between the two skis near the boots. Most of the skis were

65-70 mm wide and 1.8 to 2.0 long, although occasionally skis approximately 110 mm

wide were used. The field staff that did most of the ski penetration tests varied in mass

from 55 to 90 kg.

In Figure 6.21, these penetration measurements from 233 slabs in the Columbia

Mountains and 21 slabs in the Rocky Mountains are plotted against mean slab density and

the density estimated at 0.3 m. For the Rocky Mountain data, the average of the two

penetration measurements, PK, is not significantly correlated with mean slab density

(R2 = 0.01, p = 0.65) or with the density estimated at 0.3 m (R2 = 0.03, p = 0.49).

However, for the Columbia Mountain data, PK is significantly correlated with mean slab

density (R2 = 0.30, p < 10-4) and with the estimated density at 0.3 m (R2 = 0.50, p < 10-4).

144

Figure 6.21 Skiing penetration for mean slab density and estimated density at 0.3 m.

For this latter correlation, the linear regression of PK on ρ30 is PK = 0.55 - 0.0016 ρ30

which is used in the next section to calculate a stability index that adjusts for ski

penetration.

6.8.3 Modified Skier Stability Index

Using this stress term from Equation 6.6, a modified stability index for skier

triggering is obtained

(6.11)SK =Σφ

σxz + ∆σ xz

When the skis penetrate through the weak layer, that is PK > hSZ, SK is defined to be 0. (An

error analysis for this index is presented in Appendix B.)

The slab thickness in the start zone, hSZ, is plotted against SK in Figure 6.22 for the

same 63 persistent (square symbols) and 26 non-persistent slabs (round symbols) used to

assess SS in Section 6.7. Within the unstable range SK < 1, 76% (31/41) of the persistent

slabs and 69% (10/19) of the non-persistent slabs failed. Within the band of transitional

stability, 1 ≤ SK ≤ 1.5, 75% (6/8) of the persistent slab and none (0/5) of the non-persistent

slabs failed. In the stable range, SK > 1.5, 7% (1/14) and 0% (0/2) of the non-persistent

slabs failed.

For their unstable, transitional and stable ranges, the predictions based on SK (which

adjusts for ski penetration) and SS (which does not) are compared in Table 6.4. In the

unstable range where a higher percentage of slab failures is better, the adjustment for ski

penetration decreased the proportion of persistent slabs that failed from 81% to 76%. In

the stable range where a lower percentage of slab failures is better, the adjustment for ski

penetration reduced the proportion of persistent slabs failed from 27% to 7%.

Table 6.4 Percentage of Slabs that Failed for Skier Stability Indices

Persistent Slabs Non-Persistent Slabs

Index Index < 1 1 ≤ Index ≤ 1.5 Index > 1.5 Index < 1 1 ≤ Index ≤ 1.5 Index > 1.5

SS 81% 70% 27% 69% 13% 0%

SK 76% 75% 7% 53% 0% 0%

145

Although the proportion of false unstable predictions for persistent slabs increases

from 19% for SS (Figure 6.18) to 24% for SK (Figure 6.22), the proportion of false stable

predictions—which are critical—is reduced from 27% for SS to 7% for SK.

For non-persistent slabs, 16% fewer slabs fail in SK’s unstable range than in SS’s

unstable range. No slabs failed in the stable range of either index.

The index SK has two advantages over SS:

it is more realistic physically since it allows for ski penetration; and

the number of false stable predictions which are of greatest concern is reduced with

only a small increase in the number of false unstable predictions which have no

serious consequences.

146

Figure 6.22 Skier stability index SK for skier-tested avalanche slopes.

Since most of the points for which SK = 0 (due to PK exceeding hSZ) resulted in slab

failure, skiers appear to be efficient triggers even when the maximum penetration during

skiing exceeds the thickness of the slab.

The points for four dry slab avalanches that were skier-triggered from more than

50 m away from the avalanche are marked with a surrounding square. These remotely

triggered avalanches are discussed in Chapter 8 where it will be argued that a fracture

triggered by a skier at a localized weakness can propagate, sometimes to a nearby

avalanche slope, and release a dry slab avalanches where the slab was too stable to be

triggered by a skier. Since the snowpack properties at the trigger point can be very

different from the properties in the start zone, SK based on shear frame tests at a

well-chosen site in the start zone will occasionally yield false stable predictions. Such cases

illustrate an important limitation of any stability test done where conditions are typical of

the start zone, and emphasize that decisions about where or where not to ski should be

based on a variety of factors and not simply on one or more stability tests in the start zone.

This point is discussed further in Section 8.2 where case studies are presented for the

avalanches labelled 93-03-16 and 94-02-24 in Figure 6.22.

6.9 Predicting Skier-Triggered Avalanches on Surrounding Slopes

This section is similar to Section 6.6 which attempted to relate SN38 to natural dry

slab avalanche activity in surrounding terrain. In this section, SK is calculated for Ψ = 38o

to obtain a stability index SK38 which is related to skier-triggered avalanche activity in

surrounding terrain.

Between test days, the shear strength is interpolated linearly and the load is

calculated with Equation 6.4. However, SK38 also requires the slab height between test

days in order to calculate the shear stress due to the skier. On day j between test days i

and k, the slab height is

hj = hj-1(1-λ) + HNj (6.12)

where λ is the fractional settlement and ΗΝj is the height of the snow that fell on day j.

Since the load on day j is σv,j which can be calculated from the precipitation at a nearby

147

weather station (Eq. 6.4), HNj is simply

HNj = (σv,j - σv,j-1)/gρHN (6.13)

This leaves the density of the new snow, ρΗΝ, and settlement, λ, as the only

unknowns. For the purposes of this interpolation between test days, ρHN is taken to be

100 kg/m3 which is typical of recently deposited dry snow. Settlement depends on the

density, temperature and microstructure of each of the slab layers. However, for simplicity

and to fit the interpolated values of hj between the measured values hi and hk, fractional

settlement is calculated empirically by iteration. For the first iteration, λ = 0.03 is assumed.

If the calculated slab height on day k is more than the measured height, then λ is increased

by 0.003 for the next iteration. If the calculated slab height on day k is less than the

measured height, then λ is reduced by 0.003 for the next iteration. Iterations are stopped

when the calculated slab height on day k is within 0.02 m of the measured slab height.

During the winters of 1993-95, there were five persistent weak layers that produced

dry ski-triggered slab avalanches in the Purcell Mountains and four in the Cariboo and

Monashee Mountains near Blue River, BC. For these nine persistent weak layers, SK38 is

related to the number of skier-triggered dry slab avalanches in the following two

subsections. SK38 is assumed to have the same band of transitional stability between 1 and

1.5 as SK.

6.9.1 SK38 for Surrounding Slopes in the Purcell Mountains

On 18 January 1993 a layer of surface hoar was buried in the Purcell Mountains.

This layer was tested with shear frames at the Rocky Slope, Rocky Plot, Pygmy Slope,

and Pygmy Plot until mid-March. Although SK38 remained near 1.5 at the Pygmy Slope

until 5 February, SK38 exceeded 1.5 at the Rocky Slope and Pygmy Plot on 29 January and

2 February respectively (Figure 6.23). The last skier-triggered avalanches on the surface

hoar were reported on 27 January 1995 when SK38 was between 1 and 1.5 at these three

sites. Testing did not begin at Rocky Plot until 2 February 1995. No skier-triggered dry

slab avalanches were reported after SK38 exceeded 1.5 at any site.

On 10 February 1993, a well-developed layer of surface hoar was buried in

throughout the Columbia Mountains. Until 4 March 1993, the overlying snow was

148

generally cohesionless and

SK38 was 0 since estimated

ski penetration exceeded

the thickness of the slab

(Figure 6.24). Although

SK38 reached 1.5 on 16

March 1993 at the Rocky

Plot, it remained below 1 at

the Pygmy Plot, Pygmy

Slope and Rocky Plot until

testing was discontinued on

16 March 1993.

Intermittent skier-triggered

dry slab avalanche activity

was reported between

24 February and 9

March 1993.

However, after March

9 the instability

remained, the slab had

stiffened and was

capable of wide

propagations, and

many ski runs were

avoided because of

the buried surface

hoar. Elsewhere in the

Columbia Mountains,

there were a few

149

Figure 6.24 Skier stability trend for surface hoar layer buried10 February 1993 in the Purcell Mountains.

Figure 6.23 Skier stability trend for surface hoar layerburied 18 January 1993 in the Purcell Mountains.

reports of skier-triggered slab avalanches on the surface hoar through the remainder of

March and into April. This surface hoar layer was considered to be the most persistent

weak layer that avalanche workers had seen in many years. Certainly SK38 from Pygmy

Plot, Pygmy Slope and Rocky Slope was consistent with this perspective on the stability

of this particular weak layer.

On 6 February 1994, a layer of surface hoar was buried throughout the Columbia

Mountains. At the Pygmy Slope and Elk Plot, SK38 increased from 1 to 1.5 between

17 February and 22 February. SK38 was 1.5 when testing started at the Rocky Slope

(Figure 6.25). Most skier-triggered dry slab avalanche activity stopped on 17 March when

SK38 climbed above 1. However, there were three ski-triggered dry slab avalanches

reported to have started the surface hoar in the following 18 days. The first on 20

February was triggered by a research technician while aggressively ski-testing a steep

unsupported slope. Such isolated avalanches are expected during transitional stability. On

24 February a large slab was remotely triggered at the head of the south fork of Hume

Creek. Shear frame and

rutschblock tests were

done between the trigger

point and the slab

avalanche and are

discussed in Section 8.2.

On 7 March, a second slab

was remotely triggered in

the vicinity. Site-specific

shear frame and

rutschblock tests were

done the following day

and are discussed in

Section 8.2. As previously

mentioned, remotely

150

Figure 6.25 Skier stability trend for surface hoar layerburied on 6 February 1994 in the Purcell Mountains.

triggered dry slab avalanches may not be predicted by shear frame and rutschblock tests

where snowpack conditions are typical of the start zone. Hence, it is not surprising that

SK38 is also not a effective predictor of such slab avalanches. Except for these two remotely

triggered avalanches, no other skier-triggered dry slab avalanches were reported to have

started on the surface hoar layer when SK38 was above 1.5.

During cold clear weather between 29 December 1994 and 6 January 1995, surface

hoar grew throughout the Columbia Mountains at elevations below 1600 to 1800 m and at

a few higher elevation sites. This layer did not form at the Pygmy (2050 m) or Elk

(2250 m) sites, but did form at the Rocky Slope (1900 m) and Vermont Plot (1600 m). On

7 January, snowfall buried 3 mm surface hoar on the Rocky Slope and 10-15 mm surface

hoar at the Vermont Plot, which is located at 1550 m near the bottom of a relatively dry

valley. The snowpack and the surface hoar at the Vermont Plot were unlike those reported

in start zones. (Shear frame tests at the Vermont Plot were primarily intended for a

strength change model

that is not part of this

thesis.) SK38 from

Vermont is included in

Figure 6.26 to emphasize

that SK38 must be based

on shear frame tests

from a study site with a

snowpack that is similar

to start zones.

Except for one

remotely triggered

avalanche on 27 January,

no skier-triggered dry

slab avalanches on the

7 January surface hoar were

151

Figure 6.26 Skier stability trend for surface hoar layerburied 7 January 1995 in the Purcell Mountains.

reported after 18 January which was the last day that SK38 at the Rocky Slope was below

1.5 (Figure 6.26). The dry slab avalanche on 27 January was remotely triggered by a

group of skiers gathering at a level site to wait for a helicopter. The avalanche released

below the skiers on a short slope.

On 6 February 1995, a light snowfall buried a layer of surface hoar that had formed

in certain locations in the Purcell Mountains but not in others. In the Elk Plot, 3-5 mm

surface hoar was buried, whereas at Rocky and Pygmy Slopes, the buried surface hoar

layer could not be found. Prior to February 17, less than 0.17 m of snow covered the

surface hoar at the Elk Plot, ski penetration exceeded slab thickness and consequently SK38

was 0 (Figure 6.27). Following the snowfall that started late on 16 February and continued

until 19 February, the surface hoar stabilized rapidly and exceeded 1.5 on 23 February.

Two dry slab avalanches were skier-triggered prior to 17 February when SK38 was 0 and

none were reported after that date. In spite of the inconsistent distribution of the surface

hoar layer, SK38 was consistent with the reported avalanche activity on the surface hoar.

152

Figure 6.27 Skier stability trend for surface hoar layer buried6 February 1995 in the Purcell Mountains.

6.9.2 SK38 for Surrounding Slopes in the Cariboo and Monashee Mountains near

Blue River, BC

On 10 February 1993 a well-developed layer of surface hoar was buried throughout

the Columbia Mountains. This layer was tested with shear frames at the Mt. St. Anne

Study Plot until field work was concluded on 6 April 1993. Although no skier-triggered

dry slab avalanches were reported to have started on the layer after 1 March 1995 in this

area (Figure 6.28), many slopes remained unstable because of this layer. Also, as the slab

stiffened in March, extensive propagations were reported elsewhere in the Columbia

Mountains and many slopes were avoided where profiles revealed the presence of the

surface hoar. So, while SK38 was below 1 when avalanches were skier-triggered on this

layer, the absence of avalanches on the surface hoar layer after 1 March does not prove

that SK38 at the Mt. St. Anne study plot was too low to be of predictive value.

On 29 December 1993, a surface hoar layer of varying thickness was buried in the

Cariboo and Monashee

Mountains near Blue

River, BC. The surface

hoar crystals were

generally larger in the

Cariboo Mountains than

in the Monashee

Mountains. At the Mt.

St. Anne Study Plot in

the Cariboo Mountains,

the surface hoar crystals

were particularly large

(6-9 mm), whereas they

were 2 mm in length at

the Sam’s Study Plot in

the Monashees. No

153

Figure 6.28 Skier stability trend for surface hoar layer buried10 February 1993 in the Cariboos and Monashees near BlueRiver, BC.

skier-triggered dry slab

avalanches were

reported to have started

on this layer in the

Monashee Mountains.

The last skier-triggered

dry slab avalanche

reported to have started

on the surface hoar

occurred on 7 January

1994 when SK38 from

the Mt. St. Anne Study

Plot just exceeded 1

(Figure 6.29). At this

site, SK38 did not exceed

1.5 until 16 January. It

appears that

particularly large

surface hoar at Mt. St.

Anne was initially very weak and unusually slow to stabilize. These results from Mt. St.

Anne and Sam’s Plot emphasis the importance of initial conditions; SK38 is most effective

when the initial microstructure of the weak layer at the study site is similar to initial

microstructure of the weak layer in surrounding start zones.

On 5 February 1994, a layer of surface hoar was buried in the Cariboo and

Monashee Mountains near Blue River, BC. This layer was tested at Sam’s Plot and the

Mt. St. Anne Plot until 19 March and 22 March respectively. There were no

skier-triggered dry slab avalanches on this surface hoar layer after 18 February when SK38

at the two plots was just below 1 (Figure 6.30). SK38 exceeded 1.5 on 24 February and

4 March respectively. On 5 March in the Cariboo Mountains, a snowmobile remotely

154

Figure 6.29 Skier stability trend for the surface hoar layerburied 29 December 1993 in the Cariboo and MonasheeMountains near Blue River, BC.

triggered a dry slab avalanche

on the surface hoar when SK38

was 1.57 and 1.78 at Mt. St.

Anne and Sam’s Plots

respectively. Although SK38 is

based on the stress induced by

a skier and not a snowmobile,

the event does indicate that the

surface hoar could still be

remotely triggered.

On 7 January 1995, a

surface hoar layer that was

widespread at elevations

below 1600-1800 m was

buried in the Columbia

Mountains. This layer was

monitored with the shear

frame at the Mt. St. Anne

Study Plot until 18 February 1995. The last skier-triggered dry slab avalanche occurred on

the 25 January, the first day that SK38 exceeded 1 (Figure 6.31). SK38 exceeded 1.5, 7 days

later on 1 February 1995.

6.9.3 Summary for Skier Stability Indices

Based on the results from Figures 6.24 to 6.32, most skier-triggered dry slab

avalanches on persistent layers start when SK38 for the particular persistent weak layer at a

well-chosen study site is less than 1.5. Skier-triggered dry slab avalanches on persistent

weak layers are more common when SK38 < 1 than when SK38 < 1.5. The three

skier-triggered dry slab avalanches that occurred when SK38 > 1.5 were remotely triggered

(Figures 6.25 and 6.27) and are discussed in Chapter 8.

155

Figure 6.30 Skier stability trend for the surface hoarlayer buried 5 February 1994 in the Cariboo andMonashee Mountains near Blue River, BC.

Differences in

initial microstructure

of the persistent weak

layer between two

study sites do appear

to affect SK38. If buried

surface hoar crystals in

a particular site are

substantially larger

than at a second site,

then SK38 will tend to

be lower at the first

site and remain that

way over a period of

weeks (Figure 6.27 and

6.30). The same would

be true for start zones

in a particular valley or

at particular elevations.

6.10 Summary

Regressions for estimating the Daniels strength of common microstructures from

density are presented in Section 6.2. For those weak layers that are too thin for

density measurements, Section 6.3 shows the mean strength and variability for

Daniels strength for common microstructures by classes of hand hardness.

Sections 6.4 to 6.9 relate stability indices SN, SS and SK to avalanche activity on

slopes tested with the shear frame, and SN38 and SK38 to avalanche activity on

surrounding terrain. SN, SS are similar to S and S' developed by Föhn (1987a) except

156

Figure 6.31 Skier stability trend for the surface hoar layerburied 7 January 1995 in the Cariboo and Monashee Mountainsnear Blue River, BC.

that a microstructure-dependent normal load adjustment is applied to the shear

strength to avoid the possibility of over-estimating the stability of slopes with

persistent weak layers.

Values of SN are presented for various slopes that avalanched and did not avalanche

naturally in Section 6.5. For each of four slopes that avalanched with high values of

SN, warming or ambient temperatures near 0oC are likely explanations, showing that

SN cannot predict avalanches under such conditions.

The fact that transitional stability of SN falls well above 1 implies that a critical stress

failure criterion is not ideally suited to predicting natural avalanching. Although a

failure criterion based on critical shear strain rate would likely be a better predictor

of natural avalanching, it remains impractical since strain gauges on avalanche slopes

have not survived avalanching (Sommerfeld, 1979).

SN38 is obtained by calculating SN for a 38° inclination typical of start zones.

Although most natural avalanches of persistent slabs occurred on surrounding slopes

when SN38 was less than 2.8, SN38 ranged widely on days without reports of natural

avalanches of persistent slabs (Section 6.6). Since SN38 is similar to Σ100/ρgh which

has been used operationally for over 30 years and S35 which predicted avalanche

activity on approximately 80% of avalanche days (Jamieson and Johnston, 1993a),

the difficulty with assessing SN38 for natural avalanching of persistent slabs may lie

with reporting the failure planes for natural avalanches many of which are observed

from a distance.

Consistent with data from Föhn (1987a), SS < 1 indicates skier-triggering is likely,

1 ≤ SS ≤ 1.5 indicates marginal stability (approximately half of the tested slopes were

skier-triggered) and SS > 1.5 indicates reduced probability of skier-triggering

(Section 6.7). However, SS ignores ski penetration which is often 0.3 m in the

Columbia Mountains.

157

A practical method for estimating ski penetration based on measurements of load

and slab thickness that were already necessary for calculating SS is presented in

Section 6.8. Incorporating this estimated ski penetration into the formula for SS

results in SK which reduces the proportion of false stable predictions.

SK38 is obtained by calculating SK for a 38° inclination typical of start zones. In

Section 6.9, SK38 is shown to be better predictor of skier-triggered avalanches in

surrounding terrain than is SN38 for natural avalanches. Also, values of SK and SK38

between 1 and 1.5 correspond to transitional stability for test slopes and surrounding

terrain respectively. The indicates that the critical stress failure criterion upon which

SK and SK38 are based is effective for skier-triggered avalanches.

Differences in the initial size of surface hoar crystals between two study sites affect

stability. If surface hoar crystals at a particular site are substantially larger than at a

second site, then stability will tend to be lower at the first site and remain that way

for a period of weeks.

158

7 RUTSCHBLOCK RESULTS

7.1 Introduction

Like the shear frame stability index SK, the rutschblock test is an indicator of slab

stability for skier loading. Much as SK was assessed for avalanche slopes and for study

slopes in the previous chapter, rutschblock results from avalanche slopes are related to the

frequency of skier-triggered slabs on the tested slopes in Section 7.3, and results from safe

study slopes are related to the frequency of skier-triggered slab avalanches in surrounding

terrain Section in 7.6.

Site selection on avalanche slopes is very important since the snowpack on such

slopes is more variable than on study slopes which are chosen for their uniform snowpack.

Section 7.2 makes recommendations about site selection based on closely spaced

rutschblock tests on nine slopes which, like many avalanche slopes, include trees, buried

rocks, drifts and variations in slab thickness and slope inclination.

Since SK and rutschblock scores are both indicators of slab stability for skier

loading, there should be a relationship between them. A linear relationship is determined in

Section 7.4 based on adjacent shear frame and rutschblock tests. A breakdown in this

relationship for slopes of less than 20° is used to support a hypothesis that initial failures

on such low-angle slopes involve compression. In Section 7.5, the linear relationship and

the equation for SK are used to illustrate a method for estimating shear strength from

rutschblock scores.

7.2 Site Selection and Rutschblock Variability on Test Slopes

This section identifies some sources of variability related to terrain for the purpose

of illustrating some of the limitations of rutschblock tests and making recommendations

about selecting sites for rutschblock and other stability tests. Variability of rutschblock

scores due to changes in slope inclination has been discussed by Jamieson and Johnston

(1993b) and Schweizer and others (1995). Also, Munter (1991) has discussed similar

effects for rutschkeil tests.

159

The variability of rutschblock scores is assessed from sets of closely spaced

rutschblock tests on slopes with snowpack variability comparable to many avalanche

slopes. During the winters of 1991 and 1992, sets of 20 to 81 rutschblock tests were done

on each of nine slopes in the Cariboo or Monashee Mountains. Each slope was free of

rock outcrops and abrupt inclination changes, and, except where mentioned, the failure

plane (critical weak layer) was deeper than the operator’s skis penetrated after two jumps

on the same spot. The set of 36 tests on 6 March 1991 and the set of 81 tests on

7 April 1992 involved two operators of similar weight and the other seven sets involved

only one operator. The number of tests was limited occasionally by the amount of

undisturbed snow within the boundaries of the slope and more often by helicopter and

access logistics. Each set was completed within six hours.

On 13 February 1991, 42 rutschblocks tests were made on an east-facing slope of

Mt. St. Anne at 1900 m (Figure 7.1). The slope inclination ranged from 32° at the lowest

test positions on the

slope to 50° beside

two drifts at the top of

the slope. Except for

two tests near the

bottom of the slope,

the failures occurred in

a layer of 2-5 mm

graupel under a 0.32 to

0.51 m thick slab. At

the site of these two

tests, a weak layer of

precipitation particles

0.22 m below the

surface failed at

loading step 6.

160

Figure 7.1 Rutschblock scores from Mt. St. Anne in the CaribooMountains, east aspect, 1900 m on 13 February 1991. For mosttests with scores of six or less the slab was 0.40 to 0.50 m thickoverlying 2-5 mm graupel.

Overall, rutschblock scores ranged from 4 to 7 with most of the scores of 7 occurring at

the steep upper part of the slope where the graupel layer could not be found. It is likely

that when the rounded particles of graupel precipitated, they rolled off the steeper upper

part of the slope and were subsequently buried lower on the slope. This is consistent with

observations that weak layers of graupel are more common in gentle and moderate terrain

often used for stability tests than in start zones. In fact, graupel is not common in the

failure planes of slab avalanches although in some mountain areas it is commonly reported

in test profiles and identified as a weak layer by stability tests.

In the Cariboo Mountains on 6 March 1991, recent snowfall on top of 2-9 mm

surface hoar resulted in widespread instability. A short 25°-30° slope above a level area

was selected for repeated rutschblock tests. Twenty-six of the 36 rutschblocks failed with

score 3 (Figure 7.2) although scores

ranged from 1 to 5. Rutschblock

scores of 1 are not common but one

such result occurred in the left row

when the block slid as the side wall

was being cut. Two of the five scores

of 2 were adjacent to the test that

scored 1, suggesting a particularly

weak area of surface hoar.

Also, the four highest scores

consisting of two 4’s and two 5’s

were in the top left corner of the

slope. Surface hoar could not be

found at the sites of the top two

rutschblocks in the left row where the

blocks failed on a layer of

decomposed and fragmented

precipitation particles. It is likely that

161

Figure 7.2 Rutschblock scores from anorthwest facing slope in Miledge valley inCariboo Mountains on 6 March 1991. For mosttests, the slab was 0.50 to 0.60 m thick and slidon surface hoar.

wind near the top of the slope and possibly redirected by the tree interfered with the

formation of surface hoar, resulting in higher rutschblock scores than in the more sheltered

parts of the slope. The 4° decrease in slope angle is a less probable explanation for the

high scores in the top left portion of the slope since no comparable scores occurred at the

bottom of the slope where the inclination was also 25°. As in the Figure 7.1, the

rutschblock scores are higher and more variable near the top of the slope where wind

effects are often more evident.

On 6 April 1991, 52 rutschblock tests were made on a uniform north-facing slope on

Mt. St. Anne in the Cariboo Mountains (Figure 7.3). Slope inclinations at the sites of the

rutschblock tests ranged from 27° to 34° and did not show any consistent up-slope or

cross-slope trends. Slab thicknesses ranged from 0.44 to 0.48 m except for the two

rightmost tests in the top row where the slab thicknesses were 0.57 and 0.63 m.

Underlying this slab was a layer of 1 mm decomposed and fragmented precipitation

particles. The relatively consistent slab thicknesses and slope inclinations probably

contributed to the low variability of rutschblock scores, most of which were 3 or 4. One

exception occurred

closest to the two

trees where the score

was 6 which supports

the recommendation

that tests should be at

least 5 m from trees

(CAA, 1995). The

other exceptions

occurred at seven test

sites in the upper

three rows and are

denoted by the scores

2/3 or 2/4 in

162

Figure 7.3 Rutschblock scores from Mt. St. Anne in the CaribooMountains, north aspect, 1900 m on 6 April 1991. The slopeinclination ranged from 27° to 34°. The slab was approximately0.45 m thick. All but two blocks slid on 1 mm decomposed andfragmented precipitation particles

Figure 7.3. At these sites, a 0.17 m slab failed first on the second loading step followed by

the deeper layer on the third or fourth loading step. The scores of 3 and 4 for the deeper

weak layer are questionable since ski penetration probably increased once the 0.17 m slab

slid off the column.

On 7 January 1992, 49 rutschblock tests were made on a relatively uniform

northeast-facing slope in the Cariboo Mountains (Figure 7.4). The slope was steeper on

the left side of Figure 7.4 where slope inclinations were 34°-37° compared to 30°-33° on

the right side. Slab thicknesses ranged from 0.38 to 0.55 m. In the bottom four rows, the

scores were all 4 with

one 6. In the upper

three rows, the scores

were 3’s and 4’s with

one 5. Slab

thicknesses where the

six scores of 3

occurred were

generally thinner with

a mean of 0.42 m and

standard deviation of

0.03 m compared to

the 41 tests with

scores of 4 where

slab thicknesses were 0.48 ± 0.03 m. Also, an intermittent thin crust was reported under

the failure plane for four of the six scores of 3 and only for one test with a higher score.

This stiffer substratum may have contributed to the reduced scores by increasing the shear

stress gradient at the base of the weak layer (Schweizer, 1993). This set of rutschblock

tests also show increased variability higher on the slope.

For the test with the score of 6, no failures occurred when the operator jumped

twice near the top of the block (steps 4 and 5). After these two jumps, ski penetration was

163

Figure 7.4 Rutschblock scores from a northeast-facing slope inMiledge valley in the Cariboo Mountains on 7 January 1992.Slope inclinations ranged from 30° to 37°. The slab was 0.38 to0.55 m thick overlying 1.5 mm decomposed and fragmentedprecipitation particles.

0.42 m and the weak layer on which most tests failed was 0.10 m deeper than the skis.

However, the top 0.28 m failed when the operator stepped down towards mid-block and a

second failure occurred 0.24 m deeper on the same weak layer as other tests when the

operator “pushed” the skis down without jumping. The failure plane that was common in

other tests was deeper than average at this site—0.52 m below the surface—however

there were 18 other tests with slab thickness greater than 0.50 m so the above average

thickness was not an important factor in the above average score. Since there is no

apparent reason to doubt the score

of 6, it is clear that scores two steps

above the slope median can occur on

uniform slopes.

On 19 January 1992, 48

rutschblock tests were made on a

less uniform north-facing slope in the

Cariboo Mountains (Figure 7.5).

Slope inclinations in the shaded area

of Figure 7.5 were 30°-36° and

36°-43° in the unshaded area. In the

less steep area (shaded), 23

rutschblocks scored 4 and four tests

scored 5, whereas in the steeper area

(unshaded), two tests scored 3, eight

tests scored 4 and one test scored 5.

The slab thickness varied from 0.25

to 0.35 m over a weak layer of

1-1.5 mm decomposed and fragmented

precipitation particles. Ski penetration

after two jumps (rutschblock step 5)

ranged from 0.30 to 0.40 m. Where the

164

Figure 7.5 Rutschblock scores from anorth-facing slope in Miledge valley in theCariboo Mountains on 19 January 1992. The slabwas 0.25 to 0.35 m thick overlying a weak layerof 1-1.5 mm decomposed and fragmentedprecipitation particles. Sites where the skispenetrated the weak layer are marked “SPP”.

skis penetrated the weak layer, the result was rejected and the site marked with “SPP” in

Figure 7.5 to denote the ski penetration problem. At most of these sites, the rutschblock

failed when the operator stepped to mid-block or pushed downwards with the skis without

jumping at mid-block as part of loading step 6. However, the tests were not scored as 6’s

since the skis had penetrated the weak layer on step 5. Hence, careful monitoring of ski

penetration can avoid rutschblock scores that are high and misleading when ski

penetration approaches slab thickness. At several rutschblock sites in the steeper

(unshaded) area, the snowpack which generally exceeded 2 m in the area was only 1 m

thick over a buried rock. At such locations, the temperature gradient within the snowpack

is increased (Gray and others, 1995) and ski penetration will increase where the snowpack

has been weakened by faceting.

On 3 February 1992,

44 rutschblock tests were

made on a short north-facing

slope in the Cariboo

Mountains (Figure 7.6).

Slope inclinations ranged

from 27° to 35° with the

steepest inclinations

occurring in the middle part

of the slope. Except for four

tests in the top right part of

Figure 7.6, each rutschblock

failed on a weak layer of

2-4 mm graupel under a 0.47

to 0.62 m slab. Except for the

top right corner of the slope

where there were five scores of

7, rutschblock scores varied

165

Figure 7.6 Rutschblock scores from a north-facingslope in Miledge valley in the Cariboo Mountains on3 February 1992. Slope inclinations ranged from 27°to 35°. Except for four tests in the top right, the slabwas 0.47 to 0.62 m thick overlying a weak layer ofgraupel. * denotes rutschblocks which failed under ashallow wind slab. # denotes a rutschblock that failedin a different weak layer.

from 4 to 6 without any apparent spatial trend. The scores of 7 in the top right corner are

interesting. The weak layer of graupel was missing, presumably due to wind effect, so it

would have been a poor place for a single test to detect this particular weak layer. While it

would have been a good place to test the stability of the wind slab, many small wind slabs

can be identified by their location and smooth “chalky” appearance and a site lower on the

slope would have been a better place to test for most weak layers. This set of rutschblock

also illustrates greater variability and some higher scores near the top of the slope.

On 29 February 1992, 20 rutschblock tests were made on a northeast-facing slope in

the Cariboo Mountains. The slope inclination increased from 19° at one of the lowest tests

on the slope to 37° at the

leftmost test near the top of

the slope (Figure 7.7). All

slabs failed in the same

layer of surface hoar. The

slab was 0.40 to 0.47 m

thick except in for the top

two rutschblocks where the

slab was 0.65-0.70 m thick

due to wind-deposited

snow. For the leftmost test,

the upper 0.17 m layer

displaced on the loading

step 6 as well as the 0.42 m

slab that failed on the

surface hoar layer. The

increase in scores from 3

and 4 in mid-slope to 5 at

the bottom of the slope may

be due to the decrease in slope

166

Figure 7.7 Rutschblock scores on Mt. St. Anne in theCariboo Mountains, northeast aspect, 1900 m on29 February 1992. Except for the two tests in the drift,the slab was approximately 0.45 m thick overlying3-5 mm surface hoar.

inclination at the bottom of the slope (Jamieson and Johnston, 1993b). However, the two

scores of 6 near the top of the slope cannot be attributed to such and effect since they

occurred at the steepest sites. While the score of 6 in the top row may be due to the

thicker slab, it is unclear why the score at leftmost rutschblock test is two steps above the

slope median. This is another example of higher and more variable scores near the top of

the slope.

On 31 March 1992, 51 rutschblock tests were made on a northeast-facing slope in

the Monashee

Mountains. The slope

inclination increased

from 19°-22° for the

lowest test sites to

34°-39° for the

highest test sites

(Figure 7.8). Except

for four tests which

are marked (*) in

Figure 7.8 that failed

on a crust 0.61-0.72 m

below the surface, all

the rutschblocks failed

in a weak layer of 1 mm faceted grains and 2-3 mm surface hoar. In the lower six rows,

this weak layer was 0.39-0.57 m below the surface and in the top two rows it was

0.47-0.70 m below the surface. Rutschblock scores varied from 4 to 6 with no apparent

spatial trends. The test in the top row marked 3/5 failed 0.16 m below the surface on step

3 and then again 0.69 m below the surface on loading step 5. Rutschblock scores are not

reduced near the upper steeper part of the slope compared to their values lower on the

slope. However, a reduction in scores due to increased inclination higher on the slope may

have been obscured by the increased slab thickness in the same area. Whatever the cause,

167

Figure 7.8 Rutschblock scores on a northeast-facing slope inthe Monashee Mountains on 31 March 1992. Except for 4marked tests, the slab failed on a weak layer of 1 mm facetedgrains and 2-3 mm surface hoar.

all rutschblock scores were

within ±1 step from the median

score of 5.

On 7 April 1992, 78

rutschblock tests were done on a

northeast-facing slope in the

Monashee Mountains. A 70 kg

skier loaded the blocks on the

left side of the slope and an

80 kg skier loaded the

rutschblock on the right side of

the slope. As shown in

Figure 7.9, the slope inclination

was 29°-36° except near the top

of the slope (shaded in

Figure 7.9) where slope

inclinations were 23°-28°. The

slab thickness was 0.35 to

0.40 m in the upper less-steep part of the slope and 0.40 to 0.50 m in the lower steeper

part of the slope. Rutschblock scores range from 4 to 6. However, the four tests that

scored 6 were in the upper less-steep part of the slope suggesting that the decrease in

slope inclination near the top of the slope had more effect on the scores that the reduction

in slab thickness, which increases skier-induced stress. These scores of 6 are two steps

above the slope median. Again, the highest scores and increased variability occurred near

the top of the slope.

Good sites for rutschblocks are those with limited variability and unlikely to yield

misleading scores. It is particularly important to avoid sites that yield scores much above

average since these could contribute to an unstable slope being judged stable. As shown

by the high scores and increased variability of rutschblock scores near the top of a slope in

168

Figure 7.9 Rutschblock scores on a northeast-facingslope in the Monashee Mountains on 7 April 1992. Theblocks failed on 2-4 mm graupel under a 0.35 to0.50 m slab.

Figures 7.1, 7.2, 7.6, 7.7 and 7.9, the top of a slope is often not a good site for a

rutschblock test. In Figures 7.1, 7.2, 7.6 and 7.7, there was no decrease in slope

inclination in the upper slope where the highest scores occurred, so slope decrease was

not a factor contributing to the higher scores. Possible causes for these higher scores near

the top of a slope include:

surface hoar crystals being smaller or absent (Figure 7.2) since surface hoar growth

slows for wind speed above 2-4 m/s (Hachikubo and others, 1995) and the upper

part of a slope is often more exposed to wind than lower areas of the slope, or

graupel grains being blown off the upper part of a slope (Figure 7.6) or rolling off

the steeper part of a slope (Figure 7.1).

Alternatively, the score may be misleadingly high if a rutschblock is done where the

stress induced by the skier in the weak layer is reduced as a result of the slab being locally

thickened into a “pillow” by wind loading (Figure 7.7). Also, misleading rutschblock

results can be associated with sites near trees (Figure 7.2 and 7.3), buried rocks

(Figure 7.5) or drifts (Figure 7.7). By avoiding such sites, most rutschblock scores can be

expected to be within ±1 step of the slope median. Jamieson and Johnston (1993b)

estimate a 97% probability of rutschblock scores being within ±1 step of the slope median

on the uniform part of a slope. However, as shown in Figures 7.2 and 7.4, scores two

steps above the slope median can infrequently occur on the uniform part of a slope. The

use of other—seemingly redundant—information such as snow profiles and slope tests can

reduce the reliance on a rutschblock test and further reduce the probability of a unstable

slope being judged stable. Finally, as discussed for Figure 7.5, failure to notice that the

operator's skis have penetrated almost to, or through, the weak layer during a rutschblock

test can also contribute to over-estimating slab stability.

7.3 Rutschblocks on Skier-Tested Avalanche Slopes

Rutschblock tests were made on avalanche slopes where slab conditions were

judged typical of the start zone after the slopes were skier-tested. Occasionally, after a

slab avalanche released, no representative site could be found near the flank or crown that

169

was representative of the start zone. Usually two or more tests were done, but

occasionally there was only time or sufficient representative and undisturbed snow for one

rutschblock test. Tests were done before the slab and weak layer were judged to have

changed substantially, recognizing that snow properties change more quickly with warmer

temperatures. Except for two slopes that were tested three days after the slab avalanche,

all rutschblock tests were done within one day of the slope being skied. At slab

avalanches, only rutschblock results for the failure plane were used. On one skier-tested

slope that did not avalanche, two rutschblock results were obtained since there were two

distinct weak layers. If the skis penetrated the failure plane prior to failure of the

rutschblock, the test was rejected.

The percentage of skier-triggered slabs is plotted against 63 median rutschblock

scores from the present study in Figure 7.10 along with Föhn’s (1987b) results from a

similar but larger study in Switzerland. Non-integer median scores such as 3.5 are rounded

up. Although there are no results for median rutschblock scores of 1 in the present study,

both studies show a general decrease in the percentage of skier-triggered slab avalanches

as median rutschblock scores increase from 2 to 6. Approximately 15% of avalanche

slopes with rutschblock scores of 7 from the Swiss study were skier-triggered and Föhn

(1987b) attributes this

to difficulty selecting

sites that are safe yet

representative. In the

present study, three

of ten slabs with

median rutschblock

scores of 7 were

skier-triggered.

However, the

rutschblock tests

were done near the

170

Figure 7.10 Relative frequency of skier-triggered slabs onskier-tested avalanche slopes from Föhn (1987b) and presentstudy.

crown where slab conditions were judged typical of the start zone, yet two slopes were

triggered where the slab was much thinner, and the other was likely triggered from a spot

weaker than either the crown or the rutschblock site. These false stable results are

discussed in Chapter 8. Also, the number of these false stable results may be biased

upwards since field staff sought unusual and unexpected avalanches to determine the

limitations of rutschblock (and shear frame) tests.

In Figure 7.11, the percentage of skier-triggered slabs for the 44 slabs overlying

persistent weak layers are plotted separately from the 19 slabs overlying non-persistent

weak layers. While the number of results for non-persistent layers is limited, the frequency

of skier-triggering is clearly less than for persistent layers.

Clearly, persistent weak layers are more sensitive than non-persistent weak layers to

some difference between the skier-triggering of a slab avalanche and the skier-triggering

of a rutschblock. There are two obvious and related differences:

unlike the portion of a slab loaded by a skier, a rutschblock is not supported laterally

by the surrounding slab, and

a moving skier tests a much larger area of a start zone than a skier on a rutschblock.

Persistent weak layers may have more localized weaknesses (flaws) that are sensitive

to skier-triggering than non-persistent layers, and/or a higher percentage of fractures in

persistent weak layers

may propagate over

distances large enough

to release slab

avalanches. While the

first explanation cannot

be ruled out, field

reports strongly link

persistent weak layers

with extensive

propagations. This

171

Figure 7.11 Relative frequency of skier-triggering for persistentand non-persistent slabs on skier-tested avalanche slopes.

association may be partly due to the brittleness of persistent weak layers (Section 1.10)

and partly due to such layers remaining weak over days or weeks while the overlying slab

increases in thickness and stiffness and consequently in the strain energy capacity

necessary for extensive propagation (Jamieson and Johnston, 1992b).

If the distribution of flaws in non-persistent weak layers is similar to that of

persistent layers, then it follows that skiers are starting fractures in non-persistent weak

layers but that these fractures do not propagate over distance large enough to release slab

avalanches.

These results indicate that the microstructure of the failure plane should be

observed and reported. For example, a report of “rutschblock 4 on surface hoar” is better

for predicting skier-triggered slab avalanches than a report of “rutschblock 4”. Although

none of the 10 non-persistent slabs with median rutschblock scores of 4 to 7 were

skier-triggered (Figure 7.11), a larger study would presumably show some skier triggering

for such rutschblock results.

In Figure 7.12, the percentage of skier-triggered persistent slabs decreases from

93% for median scores of 3 or less, to 60% for median scores of 3.5 to 5, to 21% for

median scores of 5.5 to 7. For median rutschblock scores of 3 or less, 56% of

non-persistent slabs were skier-triggered and none of the slabs with median rutschblock

scores of 3.5 to 7 were skier-triggered.

The slab thicknesses at the rutschblock sites are plotted against the median

rutschblock scores in Figure 7.12. Slab thicknesses increase from 0.10-0.53 m for median

scores of 2 to 0.42-1.65 m for median scores of 7. For the three false stable results

mentioned previously (rutschblock scores of 7 near skier-triggered slabs), the thickness of

the slabs at the rutschblock sites were 1.0, 1.1 and 1.65 m. These are serious prediction

errors since such thick slabs often result in large destructive avalanches. For such thick

slabs, skiers are not effective triggers for such deeply buried weak layers because the shear

stress induced by skiers is much less than the shear stress due to the slab (Föhn, 1987a;

Figure 6.19). However, these slabs are triggered from sites with snowpack conditions

quite different from the rutschblock site, as discussed in Chapter 8.

172

7.4 Relationship Between Rutschblock Scores and SK from Adjacent

Shear Frame Tests

Rutschblock scores and shear frame stability indices such as S' (Föhn, 1987a), SS

(Jamieson and Johnston, 1993b) and SK are all indicators of slab stability for skiers. Föhn

(1987a) determined a non-linear relationship between values of the skier stability index S'

from shear frame tests and scores from adjacent rutschblock tests. Jamieson and Johnston

(1993b) regressed SS on rutschblock scores from adjacent tests and determined a linear

relationship. In this section, relationships between SK and rutschblock scores from

adjacent tests are examined for slopes with Ψ ≥ 20° and for gentler slopes with Ψ < 20°.

Based on the relationship, tests on gentle slopes (Ψ < 20°) are used to discuss a failure

mechanism for skier-triggering different from the shear failure typical of steeper slopes.

173

Figure 7.12 Slab thicknesses of persistent and non-persistent skier-tested slabs. Thepoints for some rutschblock scores are offset slightly along the abscissa for clarity.Median scores may be either integers or halves (e.g. 3.5). Data are from theColumbia and Rocky Mountains, 1992-95.

Subsequently, a relationship is established for estimating shear strength from rutschblock

scores for slopes with inclinations of 20° or more.

Between 1990 and 1995, 281 sets of shear frames tests were obtained for the failure

planes of adjacent rutschblock tests. The median rutschblock score is based on tests

usually done within 3-5 m of the shear frames tests and on the same day. Median

rutschblock scores are based on three or more rutschblock tests in 46 cases, on two

rutschblock tests in 171 cases and on one test in 64 cases.

The primary microstructure of the weak layer was classified as persistent for 207

cases, as non-persistent for 70 cases, and as unclassified for three cases. For median

rutschblock scores from 1 to 7, the mean and standard deviation of SK are given in

Table 7.1 for persistent and non-persistent microstructures. For integer-valued median

rutschblock scores, calculated values of the t-statistic and significance level, p, for a

two-tailed t-test are included in Table 7.1 to compare SK for persistent and non-persistent

microstructures. The difference between the mean values of SK for persistent and

non-persistent microstructures is only significant (p = 0.02) for a median rutschblock score

of 6 and marginally significant (p = 0.07) for a median rutschblock score of 2. Since these

two differences have the opposite sign, a systematic difference in mean values of SK

cannot be determined for persistent and non-persistent microstructures and the data are

combined with the three results for unclassified microstructures in the rightmost column of

Table 7.1.

For slopes of less than 20°, Jamieson and Johnston (1993b) reported three values of

SS for adjacent median rutschblock scores of 4, 4 and 5. However, these three values of SS

were well above the values typical for rutschblock scores of 4 and 5. To investigate this

further, 10 additional sets of shear frame and rutschblock tests were done on slopes of less

than 20° making a total of 13 sets of adjacent tests on gentle slopes. The front wall of the

rutschblock was watched closely since displacement of slabs on such shallow slopes is

often less than 20 mm. For 10 of the 13 sets of rutschblocks, fractures propagated to the

front (lower) wall for loading steps of 4, 5 or 6. The remaining three sets of tests scored 7.

This effect was analyzed using the deviations of SK from the mean for particular

174

rutschblock score, SK*, given in the rightmost column nine of Table 7.1. The normalized

deviations (SK-SK*)/SK* are plotted against slope inclination in Figure 7.13. For slope

inclination of at least 20° the normalized deviations scatter around 0 as expected.

However, for slopes of less than 20°, 10 of the 13 deviations are greater than 0. These

deviations are marked in Figure 7.13 with the median rutschblock score and the

microstructure of the weak layer. This includes 8 of the 10 results for which fractures

reached the front wall. Clearly, fractures started and propagated for rutschblock loading

steps of 4 to 6 more often than predicted by SK. Consequently, one or more of the

assumptions behind SK is inappropriate for such shallow slopes. This stability index is

based on the same assumptions as S', two of which depend on slope inclination. In the

derivation of S', Föhn (1987a) assumed that

175

Table 7.1 Skier Stability Index SK from Shear Frame Tests Adjacent to RutschblockTests for Persistent and Non-Persistent Microstructures

MedianRutsch-blockScore

PersistentMicrostructures

Non-PersistentMicrostructures

t-test All Microstructures

No. of Pairs

SK No. of Pairs

SK t p No. of Pairs

SK

Mean±SD Mean±SD Mean±SD

1 1 0.03 0 - - - 1 0.03

1.5 1 0.74 0 - - - 1 0.74

2 16 0.35±0.24 8 0.17±0.14 -1.92 0.07 24 0.29±0.23

2.5 6 0.43±0.34 1 - - - 7 0.37±0.35

3 22 0.69±0.38 10 0.52±0.32 1.28 0.21 33 0.63±0.36

3.5 5 0.65±0.59 3 0.90±0.32 - 8 0.74±0.49

4 44 0.94±0.53 17 0.91±0.57 0.19 0.85 61 0.93±0.54

4.5 4 0.87±0.16 2 1.47±0.02 - - 6 1.07±0.34

5 14 1.26±0.73 7 1.21±0.37 0.18 0.86 23 1.29±0.61

5.5 5 1.02±0.53 2 0.99±0.19 - - 7 1.01±0.44

6 35 1.64±0.89 16 1.05±0.51 2.44 0.02 51 1.45±0.83

6.5 4 1.59±0.47 2 1.94±0.46 - - 6 1.70±0.46

7 50 2.40±0.98 2 1.53±0.46 1.27 0.21 52 2.36±0.98

1. the primary fracture was in shear, and

2. the principal stress due to the slab rotates to slope-parallel prior to failure.

Although both assumptions are questionable for such shallow slopes, the first is

particularly dubious since

as the slope inclination decreases, compressive stress due to the skier will increase

and shear stress will decrease, and

the skier loading is probably responsible for fracture since snow is very sensitive to

rapid loading (Narita, 1980, 1983; Fukuzawa and Narita, 1993) .

Although in general, shear or tension is necessary for propagation, it appears that

the primary fracture for skier-triggering may be caused by compression rather than shear

on such shallow slopes. Since many of these weak layers are thin (e.g. surface hoar), this

clarifies that the compressive failures proposed by Bucher (1948) and supported by

Bradley and Bowles (1967) are not limited to thick layers of depth hoar. Skier-triggering

on shallow slopes and propagation are discussed further in Chapter 8.

176

Figure 7.13 Normalized deviations of SK from mean values for particularrutschblock scores. These deviations do not average 0 for slopes of less than20°. For these slope inclinations, the microstructure of the weak layer ismarked: SH (surface hoar), FC (faceted crystals) and DF (decomposed andfragmented precipitation particles) along with the median rutschblock score.

Excluding the adjacent shear frame and rutschblock tests on slopes of less than 20°,

means and standard deviations of SK for each median rutschblock score are given in

Table 7.2 and plotted in Figure 7.14 along with the standard error. Except for median

scores of 1.5 (for which there is only one pair), 5.5 and 6, mean values of SK increase with

increasing median rutschblock scores, indicating that SK can be estimated for particular

rutschblock scores. Jamieson and Johnston (1993b) used a regression to estimate SS from

adjacent rutschblock scores. However, the variability of SK, as indicated by the standard

deviation, increases with the mean value of SK indicating that a regression which

minimizes the sum of the squared deviations would be strongly influenced by large values

of SK which are more numerous and more variable. Fortunately, regression is not

necessary to estimate mean values of SK from median rutschblock scores. Mean values can

be read from Table 7.2 or interpolated from Figure 7.14. Interpolating between median

scores of 2 and 6.5 yields

SK = 0.31 (RB - 1) (7.1)

177

Table 7.2 Skier Stability Index SK from Shear Frame Tests Adjacentto Rutschblock Tests on Slopes of at Least 20°

MedianRutschblock Score

No. of Pairs SK

Mean St. Dev.

1 1 0.03 -

1.5 1 0.74 -

2 24 0.29 0.23

2.5 7 0.37 0.35

3 33 0.63 0.36

3.5 8 0.74 0.49

4 57 0.90 0.49

4.5 6 1.07 0.34

5 21 1.20 0.56

5.5 7 1.01 0.44

6 47 1.31 0.65

6.5 6 1.70 0.46

7 48 2.25 0.82

which is a good fit to mean values of SK except for median scores of 1.5 (for which there

is only 1 point), 5.5 and 7, which is unique since rutschblock scores have an upper bound

of 7 whereas SK has no inherent upper bound.

7.5 Estimating Daniels Strength from Rutschblock Scores

It is possible to estimate Daniels strength, Σ∞, of the failure plane from rutschblock

score since SK is a function of Σ∞ (Eq. 6.11) and SK can be estimated from the median

rutschblock score (Eq. 7.1). Combining equations 6.11 and 7.1 yields the estimated

Daniels strength

Σ∗∞ = 0.31 (RB - 1) (σXZ + ∆σXZ) - σZZ φ(Σ∞,σZZ) (7.2)

which simplifies to

Σ∗∞ = 0.31 (RB - 1) (σXZ + ∆σ'XZ) (7.3)

for persistent weak layers for which φ ≅ 0 (Section 4.11).

178

Figure 7.14 Mean, standard deviation and standard error for medianrutschblock scores from adjacent tests.

There were 208

values of SK paired with

median scores from

adjacent rutschblocks on

slopes of at least 20°.

However, ∆σXZ is not

defined for 18 cases for

which estimated skiing

penetration reached the

weak layer (h - PK ≤ 0).

Surprisingly, for the

remaining 190 pairs, Σ∗∞ is

not correlated with Σ∞ since

R2 = 0.01. However, the problem lies with ∆σ'XZ which becomes highly variable as PK

approaches h (Appendix B). For the 181 pairs (Σ∗∞, Σ∞) for which h - PK > 0.05 m, the

coefficient of determination for the correlation improves to R2 = 0.49. These 181 points

are plotted in Figure 7.15. The nine points for which h - PK ≤ 0.05 m are for relatively low

Daniels strengths (Σ∞ < 1 kPa). Four of these points are plotted with a distinct symbol in

Figure 7.15 and the remaining five lie above the graph since their estimated Daniels

strengths are between 6 and 16 kPa.

For Daniels strengths above 2 kPa, many of the estimates are too low. For Daniels

strengths below 2 kPa, estimates are proportional to measured values and most fall within

0.5 kPa of the measured value. Although rutschblock tests require less training and

specialized equipment than shear frame tests, the variability in the estimates, particularly

for strengths above 2 kPa, undermines the usefulness of estimating Daniels strength from

rutschblock scores.

179

Figure 7.15 Daniels strengths estimated from rutschblockscores plotted against measured Daniels strengths fromadjacent shear frame tests.

7.6 Relating Rutschblock Scores to Skier-Triggered Dry Slabs in

Surrounding Terrain

This section relates rutschblock scores on study slopes to skier-triggered dry slab

avalanche activity within the forecast region, typically within 15 km of the study plot. As

with the extrapolation of shear frame stability indices in Sections 6.9, this approach has

two limitations:

most of the avalanches are small since ski guides avoid slopes with weak layers

perceived to be unstable and deep enough to produce hazardous slab avalanches,

and

the failure planes of the reported avalanches are usually but not always identified.

In contrast to the shear frame stability indices, rutschblock scores cannot easily be

interpolated between test days since adjustments for increases in load due to precipitation

and decreases in slab thickness due to settlement have not been developed. Also,

rutschblock scores predicted from estimated shear strengths would not be accurate

(Section 7.5).

Initially, the relative frequencies of one or more skier-triggered dry slab avalanches

on days with results for the same failure plane from rutschblocks on study slopes were

compiled. However, there was only one match, that is, a day with one or more

skier-triggered avalanches and a rutschblock result for the same layer. The number of

matches rose to 14 when the selection was broadened to include one or more

skier-triggered dry slab avalanches that occurred within one day of the rutschblock tests.

These results are summarized by study plot and persistent weak layer in Table 7.3. For

each median rutschblock score, the fraction n/m indicates that on n of m days with a

rutschblock result, one or more skier-triggered dry slab avalanches failed in a particular

weak layer with one day of the rutschblock result for the same weak layer. However, there

were insufficient data to determine if the relative frequency of skier-triggered dry slab

avalanches decreased as the median rutschblock score increased for particular weak layers

and specific study slopes. Hence the relative frequencies are totalled by rutschblock score

for all weak layers and all study slopes in the bottom row of Table 7.3. Although this

180

totalling does not prevent the same avalanche from being counted for rutschblock results

181

Table 7.3 Relative Frequency of Skier-Triggered Dry Slab Avalanches inSurrounding Terrain within one Day of Rutschblock Tests on Study Slope

Study Slope Formation Date ofPersistent Weak Layer

Median1 Rutschblock Score on StudySlope

2 3 4 5 6 7

Rocky Study Slope, Purcells

6 December 1993 - - - 0/1 - 0/3

20 December 1993 - - - - 0/2 -

3/18 January 19932 - 1/1 - - 0/1 0/5

10 February 1993 - 1/1 1/2 - 0/1 -

19 December 1994 - - - - - 0/2

28 December 1994 - - - - - 0/2

6 February 1994 - - - - - 1/2

7 January 1995 - 2/2 - 0/1 1/1 0/3

25 January 1995 - - - 0/1 1/1 0/1

Pygmy Study Slope, Purcells

20 December 1993 - - 0/1 - 0/1 -

3/18 January 19932 - - 1/2 0/1 0/2 0/3

10 February 1993 - 1/2 - 0/1 - -

6 February 1994 - - 1/1 - - 1/2

Elk Study Slope, Purcells

6 February 1994 - - - 0/1 0/1 0/1

Mt. St. Anne, Cariboos and Monashees

5 February 1994 - - 1/1 - - 0/1

14 February 1994 0/1 - - - 0/1 -

Sam’s Slope, Cariboos and Monashees

5 February 1994 - - - - - 1/1

All Study Slopes - - - - - -

All Weak Layers 0/1 5/6 4/7 0/6 2/11 3/261 Non-integer median scores such as 3.5 are rounded up.2 Failures on surface hoar layers buried 3 January and 18 January 1993 weredifficult to distinguish since they were 5-10 mm apart after settlement.

from two different study slopes, it does show a general decrease in skier-triggered

avalanche activity as median rutschblock scores increase on study slopes (Figure 7.16).

There were at least six test days for each median rutschblock scores of 3 to 7, and except

for the fact that no skier-triggered persistent slabs were reported on six days when median

rutschblock scores averaged 5, a decrease in skier-triggered dry slab avalanche activity is

apparent in Figure 7.16. This indicates that rutschblock tests on study slopes have

predictive value for particular persistent weak layers in surrounding terrain.

7.7 Summary

Snowpack and terrain factors affecting rutschblock results are discussed in terms of

variability of rutschblock scores on nine avalanche slopes. Sites near the top of

slopes, near trees, over rocks and at pillows of wind-deposited snow sometimes

exhibit rutschblock scores and/or failure planes quite different from the remainder of

the slope. Even when avoiding such sites, rutschblock scores two steps above the

slope median are possible indicating the importance of using other sources of

182

Figure 7.16 Relative frequency of one or more skier-triggeredavalanches in surrounding terrain within one day of study-sloperutschblock results for same weak layer.

information such as avalanche activity, slope tests and profiles to confirm or raise

doubts about the results of 1 or 2 rutschblock tests.

The percentage of skier-triggered persistent slabs on skier-tested avalanche slopes

decreased from over 80% to 33% as median rutschblock scores increased from 2 to

5 (Figure 7.11). Three of nine skier-tested slopes with median scores of 7 for

persistent slabs were skier-triggered indicating that the median score from one or

two rutschblock tests is, by itself, not a completely reliable indicator of stability.

For non-persistent weak layers, no slabs were skier-triggered on 10 slopes with

median rutschblock scores above 3. The frequency of avalanching for non-persistent

slabs with rutschblock scores of 2, 3 and 4 was approximately the same as for

persistent slabs with rutschblock scores of 4, 5 and 6 respectively. Accordingly, the

interpretation of rutschblock scores should reflect the fact that persistent slabs are

more likely to be skier-triggered than non-persistent slabs with the same rutschblock

score.

A relationship between skier stability index SK and median rutschblock scores exists

on slopes with inclinations of 20° or more. However, this relationship breaks down

on slopes of less than 20° where SK usually predicts higher stability than the

rutschblock. It is hypothesized that skiers can initiate compressive fractures instead

of shear failures on slopes of less than 20°.

A method for predicting the Daniels strength based on rutschblock scores was

investigated using on the relationship between skier stability index SK and median

rutschblock scores. However, the estimates of Daniels strength are too variable to

be useful for predicting stability.

Rutschblocks on safe study slopes are shown to have predictive value for the skier

stability of persistent layers on surrounding slopes. However, as for rutschblock

tests on avalanche slopes, the particular persistent weak layer is sometimes

skier-triggered when the rutschblock score for that layer on the study slope is 7.

183

184

8 FALSE STABLE PREDICTIONS

8.1 Introduction

False stable predictions occur when a rutschblock test or shear frame stability index

indicates stability for an avalanche slope that releases under snowpack conditions similar

to those under which the tests were performed. The five skier-triggered slabs with

rutschblock scores of 6 or 7 (Figure 7.11), and/or SK values greater than 1.3

(Figure 6.22), are the focus of this chapter. (Although SK values between 1 and 1.5 are

considered transitionally stable in Chapter 7, the two slopes that avalanched with SK

values between 1.3 and 1.5 are included in this chapter since they share characteristics

with other false stable predictions.) Case studies are presented for the five avalanches. The

terrain and snowpack characteristics common to these case studies are summarized to

identify limitations of snowpack tests associated with terrain features. Also, the false stable

results are used to support an argument that some primary fractures initiated by skiers

involve compression.

8.2 Case Studies

8.2.1 Purcell Mountains, Malachite Valley, 27 January 1993

On 27 January 1993, a researcher on skis on a 5° slope felt a fracture in the

snowpack in Malachite Valley of the Purcell Mountains. The fracture propagated 100 m to

35° slope where it released a 0.8 m slab that included 0.7 m of “1 finger- to pencil-hard”

layers. The failure plane consisted of facets and surface hoar that had been buried on 18

January 1993. The slab thickness was similar at the trigger point and in the start zone.

Cracks in the bed surface, both downslope and up-slope of the crown fracture, were

apparent after the avalanche released, and are assumed to have occurred as part of the

failure process. These cracks precluded the selection of a suitable site for rutschblock

tests. Shear frame tests at a site near the crown resulted in a stability index of SK = 1.46.

185

8.2.2 Monashee Mountains, Mt. Albreda, 16 March 1993

On 16 March 1993, a slab avalanche (1.0 m thick, 30 m wide) was triggered by a

skier on a 32o north-facing moraine slope at 2100 m on Mt. Albreda in the Monashee

Mountains. At approximately six places, rocks and humps in the moraine were exposed in

the bed surface (Figure 8.1). Cracks were observed in the bed surface between the rocky

bumps in the ground surface. At the places where the snowpack was only 1 m thick prior

to the avalanche, weak depth hoar surrounded the rocks and humps. The exact trigger

point is not known but the slab was likely triggered near one of the rocks or humps

surrounded by depth hoar. The failure plane consisted of 2-3 mm rounded facets and

surface hoar from the layer of surface hoar that had been buried 18 January 1993, almost

two months earlier.

Figure 8.1 Cross-section of test site, crown fracture and substratum at slab avalancheon Mt. Albreda in the Monashee Mountains that was triggered 16 March 1993.

186

When the site was reached the next day for investigation, the most representative

undisturbed site was on the 25o slope approximately 2 m above the crown fracture. Due to

deteriorating weather, there was only time for eight shear frame tests and one rutschblock

test. (Field notes for these observations are presented in Appendix C.) At this site, which

was probably within 20 m of the trigger point, both the stability index, SK = 1.82, and the

rutschblock test, RB = 7, indicated stability. Hence, a stability test several metres away

from a localized weak spot can be misleading.

8.2.3 Purcell Mountains, Hume Valley, 24 February 1994

On a north-facing glacier at the head of the south fork of Hume Creek in the Purcell

Mountains on 24 February 1994, two researchers and a ski guide skied down gentle

terrain. They stopped on a 15-20° slope just east of the glacier near rocky outcrops where

a slab (approximately 0.2 m thick) lay on top of depth hoar. They felt a fracture in the

shallow snowpack under their skis and heard a “whumpf” sound commonly associated

with propagating fractures (snowquakes) within the snowpack (DenHartrog, 1982).

Moments later they received a radio call saying that a large slab avalanche was running

down the west-facing 35° slope approximately 400 m to the west (Figure 8.2).

The area near the crown could not be safely accessed, so a profile was observed on

the glacier approximately 150 m east of the crown where the inclination was 28° and

approximately 4 m of seasonal snowpack lay on the glacier ice (Figure 8.3). The thickness

of the slab at the profile site was 1.65 m, similar to the crown thickness that averaged an

estimated 1.5 m. The bottom 0.7 m of the slab at the profile site consisted of “pencil- to

knife-hard” layers. (Extensive fracture propagations are commonly associated with thick

slabs containing such hard and stiff layers.) The failure plane consisted of 2-6 mm facets

and surface hoar that had been buried on 6 February 1994.

Based on the shear frame tests at the 28° profile site, SK was 0.77. Calculated for the

38° slope of the start zone, SK was 0.66—remarkably low for such a thick slab and an

outlier on Figure 6.22. It is likely that the fracture had propagated through the surface

hoar layer at this profile site which was directly between the trigger point and the crown.

187

188

Figure 8.2 Remotely triggered slab avalanche in the Purcell Mountains on 24 February1994.

Figure 8.3 Cross-sections of snowpack at trigger point, profile site on propagationpath and crown for a remotely triggered slab avalanche at the head the south fork ofHume Valley in the Purcell Mountains on 24 February 1994.

Approximately 1 h had elapsed between the fracture propagation and the shear frame and

rutschblock tests. During this time, the fractured surface hoar layer may have partly

rebonded under the weight of the slab. Although this questionable value of SK suggests

instability, the rutschblock result indicated stability. Even when three people without skis

jumped on the rutschblock simultaneously, the block did not fail. This rutschblock score of

7 is not surprising since—for such deep weak layers—the stresses induced by persons on

foot are small compared to the stress induced by such a thick slab (Föhn 1987a;

Figure 6.19), a limitation of the rutschblock test identified in Section 7.3.

It is very likely that the fracture was initiated by the skiers in the depth hoar near the

rocky outcrops, propagated through the surface hoar layer in a snowpack that could not

be triggered by skiers and released a large slab avalanche when it reached a slope steep

enough to avalanche.

8.2.4 Purcell Mountains, Hume Valley, 8 March 1994

On 8 March 1994, a second slab avalanche was remotely triggered by skiers at the

head of the south fork of Hume Creek in the Purcell Mountains. Skiers in gentle terrain on

a west-facing slope at 2415 m initiated a fracture within the snowpack that propagated

60 m to a 35° slope where it released a 1.0-m-thick slab avalanche. A profile on a 14°

slope at the estimated trigger point revealed a 1.0-m-thick slab of which 0.7 m consisted

of “pencil-hard” layers overlying a layer of 8 mm surface hoar that had been buried on 6

February 1994. A profile near the crown was similar, showing a 0.9 m slab that included

0.75 m of “pencil-hard” layers overlying the failure plane of 1-8 mm facets and surface

hoar. This is another example of a fracture that initiated in low angle terrain and

propagated to a slope steep enough to slide where it released a slab avalanche.

There was no suitable site for rutschblock tests near the crown. However, shear

frame tests at the profile site resulted in a SK value of 1.37 for the 35° slope of the start

zone.

189

8.2.5 Observation Peak, Rocky Mountains, 9 February 1995

A skier ascending a southeast-facing 20° slope at 2400 m on Observation Peak in

the Rocky Mountains triggered a fracture that propagated 300 m to a 34° slope where it

released a 100 m wide slab avalanche. The slab was approximately 0.5 m thick at the

trigger point and 1.1 m thick at the crown where it included 1.0 m of “1-finger and

pencil-hard” layers. At a 34° site near the crown, a rutschblock test indicated stability

since the block did not slide after a skier jumped several times on the block (RB = 7).

Hence, the slab was released by a propagating fracture triggered where the slab was less

stable, but the slab remaining in the start zone could not be skier-triggered.

8.3 Characteristics Associated with False Stable Predictions

Selected snowpack and terrain characteristics for these five case studies are

summarized in Table 8.1. Certain characteristics are common to these false stable

predictions:

In every case, the failure plane consisted of a persistent microstructure.

In four of the five cases, the trigger point was more than 50 m from the avalanche,

so slab conditions at the trigger point may have been very different from the start

zone and the site of the shear frame and rutschblock tests. In each of these cases of

remote triggering, the avalanche was triggered from a slope too shallow for a dry

slab to avalanche (< 25o).

In two of the four cases, the slab at the trigger point was much thinner than at the

crown.

In at least two cases, the fracture initiated in depth hoar near rocks.

In two of five cases, obvious cracks extended through the bed surface.

8.4 Remote Triggering and Transitional Stability for SK

During the winters of 1992-93 to 1994-95, shear frame and/or rutschblock tests

were made at 95 skier-tested avalanche slopes. Of the five slopes that produced false

stable results, four were triggered from more than 50 m away from the resulting

190

avalanche. Assuming that propagating fractures (snowquakes) can advance through weak

layers in which a skier could not start a fracture, it follows that skiers in the start zone

where SK > 1.3 or RB = 7 may have been unable to trigger the slabs that were remotely

triggered. Since the only skier-triggered slabs for which SK > 1.3 were remotely triggered,

the results summarized in Figure 6.22 imply transitional stability for 1 < SK < 1.3.

However, until more data are available to refine the band of transitional stability, an upper

limit of 1.5 appears to provides a reasonable margin of safety.

The avalanche on Mt. Albreda that was not triggered remotely is important because

it illustrates that shear frame and rutschblock tests done where conditions appear typical of

the start zone can incorrectly indicate stability.

191

Table 8.1 False Stable Predictions

Date Location Predictor

SlopeInclination

(o)

SlabThickness

(m) Comments

Trig.Point

StartZone

Trig.Point

StartZone

93-01-27 Purcell Mtns.,Malachite

ValleySK = 1.46 5 35 0.8 0.8

Remote trig. 100 m.Cracks through bed

surface

93-3-17 MonasheeMtns.,

Mt. AlbredaSK = 1.82RB = 7

32 32 1.0 1.0Depth hoar around

rocks at trigger point.Crack through bed surf.

94-02-24 Purcell Mtns.,Hume Creek

RB = 7SK = 0.66 20 38 0.2 1.5

Depth hoar aroundrocks at trigger point.Remote trig. 400 m.

94-03-08 Purcell Mtns.,Hume Creek

SK = 1.37 14 35 1.0 1.0 Remote trigger.60 m

95-02-09 Rocky Mtns.,Observation

PeakRB = 7 20 34 0.5 1.1 Remote trigger 300 m

8.5 An Alternative Failure Mode for Primary Fractures

The skier stability index, SK, is a refined version of Föhn’s (1987a) S' for skiers.

Both indices are based on the ratio of shear strength to shear stress. The success of these

indices for predicting slab stability for skiers (Figure 6.22) is proof that failures for most

skier-triggered slabs begin with a shear failure within a thin weak layer.

However, important exceptions to shear failure may occur on slopes of less than 20°

inclination. Section 7.4 shows that SK, which is based on shear failure, overestimates the

stability of rutschblocks on such slopes. Presumably, the initial failures on slopes of less

than 20° that SK fails to predict involve compression. Further, all four of the remotely

triggered slabs for which SK and/or RB incorrectly indicated stability were triggered on

slopes of 20° or less. Even without specifying the failure mode on shallow slopes, SK

cannot be expected to predict skier-triggered avalanches on such shallow slopes because it

cannot reliably predict rutschblock failures on such shallow slopes. The rutschblock test,

which is not restricted to initial shear failure, may prove useful for assessing the potential

of snowpacks on such shallow slopes for initiating fractures. Certainly, the rutschblock

test is capable of identifying weak layers on such shallow slopes, regardless of the failure

mode.

Bed surface cracks were reported in two of the four false stable results (Table 8.1).

Although the bed surface was not photographed at either site, Figure 8.4 shows bed

surface cracks at an avalanche at Whistler Mountain in February 1979. At the remotely

triggered avalanche in the Malachite Valley on 27 January 1993, cracks through the bed

surface were found well above the crown and precluded the selection of a suitable site for

rutschblock tests. If such cracking were simply a consequence of the avalanche then it

would likely be reported at more of the 52 investigated avalanches. However, bed surface

cracks were only observed at two avalanche sites, both of which gave false stable results.

The failure of SK to predict instability at these sites, and particularly at the avalanche on

Mt. Albreda which was not triggered remotely, could be explained if the initial failure

involved the cracks and not shear failure in the weak plane along which the fracture

subsequently propagated. While this argument is far from conclusive, it and SK’s

192

over-estimation of

rutschblock stability on

shallow slopes (Section

7.4) suggest that not all

cases of skier-triggering

begin with shear failure

of a weak layer. The

bed surface cracking is

consistent with a

primary compressive

fracture at the base of

the snowpack or within

thick depth hoar layers

as proposed by Bucher

(1948), Bradley (1966),

Bradley and Bowles

(1967) and Schweizer

(1991). However,

primary compressive

fractures may not be

limited to thick weak

layers. SK’s

over-estimation of

rutschblock stability on

shallow slopes with weak layers of surface hoar (Figure 7.13) suggests that primary

compressive failures can occur within relatively thin layers.

193

Figure 8.4 Cracks in bed surface at Whistler Mountain,February 1979. Such cracks are believed to occur during slabfailure. (C. Stethem photo)

8.6 Summary

Sections 6.8 and 7.3 show that the skier stability index, SK, and rutschblock scores

based on tests where conditions are judged typical of start zones (Figure 6.22 and 7.12)

can predict the skier stability of most slopes. However, as the case studies illustrate,

stability tests done where snowpack conditions are typical of start zone are occasionally

misleading and cannot predict avalanches triggered at localized weaknesses—sometimes

with dimensions of only a few metres—or remotely from sites with a less stable snowpack

than the start zone. This represents an important limitation of stability tests since it is

impractical to test all potential trigger points associated with a locally thin snowpack or

with humps, rocks, trees or bushes under the snowpack that are within a few hundred

metres of start zones. Stability tests and profiles are presently interpreted together with:

a general awareness of the snow distribution,

knowledge of mesoscale stability trends based on weather, study site and avalanche

observations, and

familiarity with the terrain. Clearly, the character of the ground-snow interface

within a few hundred metres of the start zone is a relevant terrain consideration.

Although this idea is not new, the case studies of false stable predictions confirm its

importance.

Localized weaknesses should be suspected wherever the ground surface is

particularly uneven, as is common on moraines. This is particularly important when stiff

slabs overlie persistent weak layers—a combination capable of extensive propagation. The

more extensive the propagation, the more likely the fracture will reach a slope steep

enough to avalanche. However, although a weak layer and a stiff slab are required for

propagation, there is presently no practical snowpack test that indicates whether local

fractures that can start near rocks, bushes or thin snowpack areas, etc. will propagate over

tens or hundreds of metres, or not at all.

194

9 APPLICATIONS OF SHEAR FRAME STABILITYINDICES TO AVALANCHE FORECASTING

9.1 Introduction

In conventional avalanche forecasting, the forecaster’s experience is used to

anticipate avalanche activity based on observations of weather, snowpack and past

avalanches (LaChapelle, 1980; Buser and others, 1985). Some forecasting operations

presently use shear frame stability indices and rutschblock tests along with other weather,

snowpack and avalanche observations to make decisions. Since 1990, rutschblock tests

have been adopted to varying degrees by backcountry avalanche safety programs in

Canada. Although shear frame stability indices have been used for forecasting natural

avalanches of storm snow and closures for timing explosive control at the highways

through Rogers Pass since 1963 (D. Skjönsberg, personal communication) and through

Kootenay Pass since 1980 (J. Tweedy, personal communication), such indices are

presently not used by backcountry avalanche forecasting programs in Canada where

skier-triggered avalanches are the greatest concern.

This chapter attempts to determine if extrapolated shear frame stability indices could

improve backcountry avalanche forecasting of persistent dry slabs which are the cause of

most backcountry fatalities (Jamieson and Johnston, 1992a). The approach is to compare

the number of days correctly forecast using SN38 and SK38 as well as conventional

measurements with the number of days correctly forecast using only conventional

measurements. A limitation of this approach results from the selection of conventional

measurements. Quantitative meteorological measurements such as air temperature,

precipitation and wind speed taken daily at fixed sites and previous avalanche activity are

used. However, snowpack tests such as shovel tests, compression tests, ski tests and

profiles are excluded since they are done intermittently and at varying locations. Although

backcountry forecasters consider such tests to be important, they are difficult to assess

systematically.

195

Sections 9.2 and 9.4 present correlations of individual forecasting variables such as

previous avalanche activity, air temperature, precipitation, wind speed and shear frame

stability indices with natural and skier-triggered avalanche activity, respectively. Sections

9.3 and 9.5 develop simple multivariate forecasting models for natural and skier-triggered

avalanches respectively, and compare the performance of these models when shear frame

stability indices are included and excluded.

The analysis in each section is repeated for the Purcell Mountains near Bobby Burns

Lodge and for the Cariboo and Monashee Mountains near Blue River, BC.

Measurements for the meteorological variables are taken in the morning before

skiing terrain is selected for the day. Interpolated values of the shear frame stability

indices, SN38 and SK38, are used between days that persistent weak layers were tested with

the shear frame. All variables are related to avalanche activity for the same day. While the

date of occurrence is accurately recorded for skier-triggered avalanches, the occurrence

date of natural avalanches is often estimated. Avalanches that are estimated to have

occurred during the night prior to morning weather observation are usually recorded as

having occurred on the previous day. This is optimal since it relates avalanche activity to

daily weather measurements such as 24 h maximum air temperature, 24 h snowfall, etc.

Various measures of avalanche activity are possible. The daily total number of

avalanches (with a maximum of 10) was used in Chapters 6 and 7. McClung and Tweedy

(1994) summed the size classes for all reported avalanches. Davis and Elder (1995)

compared various measures of avalanche activity including number of avalanches, sum of

sizes of all avalanches, as well as the size of the largest avalanche on a given day, and

found that the ranked order of forecasting variables was the same for each measure of

avalanche activity. This implies that the assessment of a particular variable’s predictive

value—which is an objective of this chapter—is relatively insensitive to the measure of

avalanche activity. For the present data set, Spearman rank correlations between the daily

number of persistent dry slab avalanches and the daily maximum size class are 0.999 for

natural avalanches and 0.997 for skier-triggered avalanches for 356 days in the Cariboo

and Monashee Mountains. Similarly, the correlations are 0.999 for both natural and

196

skier-triggered avalanches for 295 days in the Purcell Mountains. Such correlations may

seem surprising since large avalanches are much less common than small avalanches.

However, since no persistent dry slab avalanches are reported on many days, the various

measures of avalanche activity primarily distinguish between days with avalanches and

days without avalanches.

In subsequent analyses, the daily maximum size class of natural or skier-triggered

avalanches involving a persistent slab, MxN or MxS, respectively, is used as the measure of

avalanche activity in part because the size of expected avalanches—especially

skier-triggered avalanches—affects backcountry decisions more than the number of

avalanches. For example, when class 1 avalanches (not large enough to injure a person)

are expected, ski guides will intentionally trigger many slabs to stabilize slopes and, in

many cases, remove the weak layer before additional snowfall builds a thicker and more

destructive slab. In contrast, slopes are generally avoided on which a class 2 slab

avalanche (large enough to injure, bury or kill a person) might occur.

Since the forecasting model described in the next section tends not to predict levels

that rarely occur, the number of levels of avalanche activity should be reduced for these

analyses. For this reason, half-sizes of avalanches (CAA, 1995; Table 3.2) are rounded up

to the nearest integer, and MxN and MxS are assigned a value of 3 on the rare days with a

persistent slab avalanche larger than class 3. As a result of this reduction, MxN and MxS

only take on values of 0, 1, 2 or 3, levels which are adequate for practical

decision-making.

The following analyses use the shear frame stability indices for the persistent weak

layers discussed in Chapter 6, excluding those from the Rocky Mountains for which

insufficient data were available. In the Purcell Mountains near Bobby Burns Lodge, the

tested persistent layers were buried 19 January 1993, 10 February 1993, 6 February 1994,

7 January 1995 and 6 February 1995. In the Cariboo and Monashee Mountains near Blue

River, BC, the tested persistent weak layers were buried 10 February 1993, 29 December

1993, 5 February 1994 and 7 January 1995. To obtain one daily value of SN38 and of SK38

for each of the two forecast areas, values for various persistent weak layers from different

197

study sites within a forecast area are averaged for each day. Averaging across study sites

reduces the effect of an unusually weak (or strong) persistent layer at a particular study

site such as occurred for the surface hoar layer buried 10 February 1993 at Mt. St. Anne

(Section 6.9.2).

9.2 Forecasting Variables for Natural Avalanches of Persistent Slabs

This section presents relationships between the size class of the largest natural

persistent dry slab avalanche reported on a particular day, MxN, and

the extrapolated stability index, SN38, for natural avalanches of persistent slabs,

local meteorological measurements available to the forecasters and ski guides on the

morning of the day,

MN1 which is the size class of the largest natural persistent slab avalanche reported

for the previous day, and

MN2 which is the sum of the size classes of the largest natural persistent slab

reported for the two previous days.

The relationship between two variables may be either monotonic (increasing or

decreasing) or non-monotonic, or there may be no discernible relationship. For example, a

monotonic (increasing) relation would exist between wind speed and avalanche activity if

increased wind speed was associated with increased avalanche activity. If more activity

were associated with moderate winds than with light or strong winds, then the relationship

would be non-monotonic. Monotonic relationships are assessed with Spearman rank

correlations which are suited to ordinal data. Non-monotonic relationships are assessed

with box graphs.

Spearman rank correlation coefficients, R, are presented in Table 9.1 for MxN’s

relationship to common meteorological variables as well as to MN1, MN2 and SN38 for

observations the winters of 1992-93, 1993-94 and 1994-95. These meteorological

measurements include total height of the snowpack, HS, height of 24 h snowfall, HN,

accumulated snowfall during a storm, HST, and foot penetration, PF, which can also be

considered snowpack measurements. In the Purcell Mountains, the meteorological

198

199

Table 9.1 Spearman Rank Correlations Between Forecasting Variables and the DailyMaximum Size of Natural Avalanches Involving Persistent Slabs, MxN

Forecasting VariablesPurcell Mountains1 Cariboo and

Monashee Mtns.2

N R p N R p

Max. Nat. Av. on Previous Day (MN1) 294 0.03 0.61 355 0.38 <10-6

Sum of Max. Nat. Av. Previous 2 Days(MN2)

293 0.21 <10-4 354 0.47 <10-6

Barometric Pressure (BP) 261 0.04 0.51 353 -0.12 0.02

Air Temperature (Ta) 268 0.03 0.62 327 -0.01 0.87

24 hr Min. Air Temperature (Tmin) 267 0.06 0.35 332 -0.01 0.85

24 h Max. Air Temperature (Tmax) 267 0.10 0.12 334 -0.10 0.07

Wind Speed (WS) 264 0.03 0.59 258 0.05 0.38

Wind Direction (WD) 75 -0.03 0.77 258 0.10 0.13

Height of 24 h Snow (HN) 264 -0.06 0.37 - - -

Height of Storm Snow3 (HST) 255 -0.03 0.61 - - -

Water Equiv. of 24 h Precip.4 (HNW) - - - 323 0.15 0.01

Water Equiv. of Storm Precip.5 (HSTW) - - - 323 0.16 10-2

Height of Snowpack (HS) 268 0.02 0.77 356 0.06 0.26

Foot Penetration (PF) 263 0.03 0.60 - - -

Natural Stability Index6 (SN38) 138 -0.21 0.01 168 -0.01 0.911 All variables except for SN38 based on manual measurements at Bobby Burns Lodge,1370 m at approximately 0630 h. Wind speed estimated as calm, low, moderate orstrong and converted to 0, 15, 35 or 50 km/h. Wind Direction estimated using 8cardinal directions.2 All variables except for barometric pressure measured automatically at Mt. St. Anne,1900 m. Air temperature, wind speed and wind direction are averaged between 0400and 0500 h. 3 Reset to 0 after precipitation has stopped and useful settlement observations wereobtained.4 Estimated from HN when precipitation gauge not working.5 Cumulated HNW. Reset to 0 after 24 h period ending at 0500 with less than 0.3 mmprecipitation.6 Mean of measured and/or interpolated values from various study sites and persistentweak layers.7 R values marked in bold are significant at the 0.05 level.

measurements were obtained in the morning from manual weather observations at Bobby

Burns Lodge (1370 m). For the Cariboo and Monashee Mountains near Blue River, BC,

meteorological measurements, other than barometric pressure were obtained from an

automatic weather station (1900 m) on Mt. St. Anne in the Cariboo Mountains.

Barometric pressure was obtained from the weather station at the Blue River airport

(678 m) and adjusted to the equivalent pressure at sea level. Storm and 24 h precipitation

measurements are HSTW and HNW in mm from a gauge that melts the precipitation to

determine the equivalent amount of water. Differences in the observations recorded at

each area reflect differences in equipment and operational practices at the two areas.

Of the variables given in Table 9.1 for the Purcell Mountains, only MN2, and SN38

are significantly correlated (p < 0.05) with MxN. The positive correlation of MN2 with

MxN is consistent with the accepted use of recent avalanche activity as a predictor of

expected avalanche activity. The negative correlation between SN38 and MxN implies that

MxN tends to increase as the natural stability index, SN38, decreases. On the many days that

no natural avalanches are reported for persistent slabs (MxN = 0), high values of SN38 are

implied. The lack of correlations with the meteorological forecasting variables other than

Tmax may be due to the difference in weather between the start zones, typically 1700 m to

2700 m, and the Bobby Burns Lodge which is at 1370 m in a relatively dry valley, or with

difficulty identifying persistent failure planes for natural avalanches—most of which are

observed from a distance. Difficulty identifying the failure plane for natural avalanches

may partly explain why persistent avalanche activity on the previous two days (MN2)

correlates better than activity on the previous day (MN1).

In the Cariboo and Monashee Mountains near Blue River, BC, MN1, MN2, PB,

HNW and HSTW are significantly correlated with MxN (p < 0.05). The strong positive

correlations of MxN with MN1 and with MN2 indicate the importance of previous

avalanche activity as a predictor of present natural avalanche activity. The negative

correlation with barometric pressure implies that natural avalanches of persistent slabs are

more common on days with low barometric pressure than for higher pressure. The

positive correlations with 24 h precipitation and storm precipitation indicate that persistent

200

avalanches are more common for higher values of precipitation than for lower values. The

correlation of MxN with SN38 is not significant which is consistent with the discussions in

Section 6.6.2.

Non-monotonic relationships between MxN and most forecasting variables from

Table 9.1 are assessed with box graphs in Figure 9.1. Foot penetration is excluded from

Figure 9.1 since the usual interpretation—deeper foot penetration is associated with larger

201

Figure 9.1 Box plots of the daily maximum size of natural avalanche involving apersistent slab against various forecasting variables showing median (small rectangle),lower and upper quartiles (box) and minima and maxima (whiskers). Boxes for thePurcell Mountains near Bobby Burns Lodge are unshaded and boxes for the Cariboosand Monashee Mountains near Blue River, BC are shaded. Precipitation values areheight of snowfall in cm for the Purcell Mountains and height of melted precipitation in

avalanches—implies a monotonic relation which was tested with the correlation in

Table 9.1. Air temperature, Ta, and minimum 24 h air temperature, Tmin, are excluded

since they exhibit the same general trend as Tmax which is plotted in Figure 9.1. MN1 is

excluded since MN2 yields stronger correlations with MxN in both forecast areas. Height

of 24 h snow, HN, and height of storm snow, HST (cm), from the Purcell Mountains are

plotted along with their respective water equivalents, HNW and HSTW (mm), from the

Cariboo and Monashee Mountains.

Non-monotonic relations apparent in Figure 9.1 are used to select variables for a

multivariate model in the next section. For each forecasting variable and each value of

MxN, the median values are plotted as small rectangles and the boxes which extend from

the 25th to 75th percentile include the middle half of the data. Substantial shifts in the

median values or boxes for different values of MxN suggest that the variable may have

predictive value.

In addition to MN2, PB, HNW and HSTW which correlate with MxN in the Cariboo

and Monashee Mountains (Table 9.1), Tmax, WS, WD, and HS, also show promise of

being useful predictors (Figure 9.1).

In the Purcell Mountains, HS and Tmax show promise as predictors. Three variables

which are useful predictors for other forecasting models (e.g. Buser and others, 1987;

Davis and others, 1993; McClung and Tweedy, 1994), namely, wind speed, WS, height of

24 h snowfall, HN, and height of storm snow, HST, do not appear to be promising

predictors of MxN, probably because they are observed at Bobby Burns Lodge where

snow and weather conditions are quite different from start zones in the Purcell Mountains.

9.3 A Multivariate Forecasting Model for Natural Avalanches Involving

Persistent Slabs

9.3.1 Selection of Model

The objective of this section is to determine if the inclusion of SN38 in a multivariate

forecasting model improves the performance of the model. Although nearest neighbours

models have advantages (Section 1.8) for forecasting, such models are not suited to

202

assessing the importance of a particular variable since they require that the variables be

weighted heuristically (Buser and others, 1985). Classification tree models are used in

preference to discriminant analysis for this assessment of SN38, since such models do not

require normalizing transformations, allow for complex interactions between the predictor

variables, are sensitive to non-monotonic relations between the predictor variables and the

response variables, and allow a categorical response variable with more than two levels for

avalanche activity (Davis and Elder, 1995).

Classification trees recursively split the data into two groups using various

partitioning rules. Fortunately, the resulting trees tend to reflect structure in the data and

are not strongly affected by the choice of partitioning rule (Breiman and others, 1984,

p. 94). Although a partitioning rule based simply on the number of cases (days)

misclassified is tempting, Breiman and others (1984, p. 94-98) prove otherwise. The

partitioning rule of the S-Plus software that was used for these analyses is based on

deviance which is a measure of a lack-of-fit of an observation to the data used to

construct a particular node. Consider a day with MxN = 1 directed to a particular node of

the tree. If all the data used to construct the node had MxN = 1, then the deviance for the

particular day would be zero at that node. If only 60% of the observations used to

construct the node had MxN = 1 then the deviance would be greater than zero according

to the log-likelihood formula for deviance (Chambers and Hastie, 1992, p. 412-414).

For the present data, each forecasting variable consists of ordinal (or interval) values

which allows each variable, Xi, to be split using a critical value, Xic. At each split, each

critical value, Xic, between sorted values of Xi is tried for each variable to find the split

Xi < Xic that partitions the data into subsets with minimal deviance. The same splitting rule

is then applied to each subset. The fact that the same forecasting variable can be used

recursively allows complex patterns in the data to be detected.

Potentially, sets could be split until there is only one datum in each subset. However,

while the initial splits reflect structure and grouping of the data (which are important),

splitting into very small subsets results in fitting a tree to individual data points (which is

not relevant to most problems). For the present application, splitting was stopped when

203

there were five or less days in a subset, or the set consisted of points with a single value of

MxN, indicating that five or more days with the same maximum size of dry natural

avalanches had been grouped together. Subsets that are not subdivided further are called

terminal nodes or leaves.

The measure of the lack-of-fit of a particular tree to a compatible data set is the

residual mean deviance defined as the deviance summed over all the observations divided

by the degrees of freedom of the tree (number of cases minus the number of terminal

nodes) (Statsci, 1994, p. 12.10). Hence, the better a tree fits a data set, the less the

residual mean deviance. Although the misclassification rate is the obvious measure of

lack-of-fit, it ignores the fact that the probability of a particular value of the response

variable, MxN at a node usually falls between 0 and 1 (Breiman and others, 1984,

p. 94-98). Subsequent analyses present the lack-of-fit in terms of both the misclassification

rate and the residual mean deviance.

9.3.2 Purcell Mountains

All data-based multivariate models including classification trees require large data

sets (Davis and Elder, 1995). However, the size of the data set (number of days) tends to

decrease as the number of variables increases since different variables often have missing

values on different days. For example, including SN38 in a model excludes those days that

neither measured nor interpolated values of SN38 were available.

During the three winters that persistent weak layers were monitored in the Purcell

Mountains, measured or interpolated values of SN38 were obtained for persistent weak

layers from at least one study site between 26 January 1993 and 16 March 1993, between

13 February 1994 and 21 March 1994 and between 12 January 1995 and 3 March 1995

for a total of 138 days. The daily avalanche activity for the weak layers that produced one

or more dry slab avalanches is described in Sections 6.6 and 6.9.

For the Purcell Mountains, the rank correlations in Table 9.1 show that MN2, Tmax

and SN38 exhibit significant monotonic relationships with MxN, and the box graphs

(Figure 9.1) for PB, HS and possibly storm precipitation, HST, show non-monotonic

204

relationships to MxN. Eliminating the days for which any of these six variables had missing

values yields a data set of 12 days with persistent slab avalanches (MxN > 0) and 121 days

without such avalanches (MxN = 0).

When the classification tree algorithm is applied to the variables MN2, HS, HST,

Tmax and PB with SN38 excluded, a tree (Model N-P-E in Table 9.2, Figure 9.2) results

which misclassifies 12 days and has a deviance of 0.44. Removing the parts of the tree that

do not reduce the misclassification rate (circled in Figure 9.2) reduces the number of

terminal nodes to four and increases the residual mean deviance to 0.56. Following the

splits (“decisions”) in Figure 9.2, MxN = 2 is “predicted” when 0.80 ≤ HS < 1.15 m and

MN2 ≥ 3. (Of course, such predictions reflect the limited data set and may not fit

expectations based on intuition and determinism.)

Including SN38 with MN2, HS, HST, Tmax and PB for the same set of 133 days

results in Model N-P-I (Figure 9.3) which has a reduced residual mean deviance of 0.37

and a reduced misclassification rate of 11/133. Nine of 12 avalanche days are misclassified

205

Figure 9.2 Classification tree for daily maximum size of natural avalanches of persistentslabs in the Purcell Mountains using forecasting variables but excluding SN38. Data arefrom the winters of 1992-93 to 1994-95. For each split, the left branch denotes days forwhich the “less than” criterion is true.

206

Figure 9.3 Classification tree for the daily maximum size of natural avalanches ofpersistent slabs in the Purcell Mountains using forecasting variables including SN38.Data are from the winters of 1992-93 to 1994-95. For each split, the left branchdenotes days for which the “less than” criterion is true.

Table 9.2 Classification Trees for Daily Maximum Size of Natural Avalanches ofPersistent Slabs in the Purcell Mountains

ModelName Forecasting Variables1

Cost-Complexity

Factor

No. ofTerminalNodes

ResidualMean

Deviance

Mis-classifi-cationrate

N-P-E HS, HST, Tmax, PB, MN2 0 14 0.44 12/133

N-P-I SN38, HS, HST, Tmax, PB, MN2 0 11 0.37 11/133

N-P-1 SN38, HS, HST, Tmax, PB, MN2 1 9 0.38 11/133

N-P-2 SN38, HS, HST, Tmax, PB, MN2 2 7 0.43 11/133

N-P-3 SN38, HS, HST, Tmax, PB, MN2 3 7 0.43 11/133

N-P-4 SN38, HS, HST, Tmax, PB, MN2 4 5 0.48 11/133

N-P-5 SN38, HS, HST, Tmax, PB, MN2 5 4 0.53 11/133

N-P-6 SN38, HS, HST, Tmax, PB, MN2 6 4 0.53 11/133

N-P-7 SN38, HS, HST, Tmax, PB, MN2 7 2 0.64 12/1331 Variables marked in bold are selected by the classification tree algorithm from thoselisted and used to build the tree.

and two of 121 non-avalanche days are misclassified as shown in Table 9.3. So, including

SN38 results in one more avalanche day being correctly classified. Removing the parts of the

tree that do not reduce the misclassification rate (circled in Figure 9.3) reduces the number

of terminal nodes to four and increases the residual mean deviance to 0.53.

To determine which of the six variables are most effective at reducing the residual

mean deviance, D, the tree model can be simplified by increasing the cost-complexity

factor, k, and removing the least important subtrees T' with cost-complexity (Statsci,

1994, p. 12.17) defined as

DK(T') = D(T') + k · number of terminal nodes of subtree T'. (9.1)

By increasing the cost-complexity factor, k, from 1 to 7, the residual mean deviance and

misclassification rate increase, the number of terminal nodes decreases and the most

important variables are retained at each step (Table 9.2). By this technique, the forecasting

variables for MxN in the Purcell Mountains are, in decreasing predictive value, SN38, HS,

HST, Tmax, PB and MN2. Of these variables, only SN38 and Tmax showed significant

monotonic relationships to MxN in Table 9.1. The second most important variable, HS, did

not exhibit a significant correlation with MxN (Table 9.1) but did show a non-monotonic

207

Table 9.3 Contingency Table for Natural Slab Avalanches Involving Persistent Slabs inthe Purcell Mountains

ModelledSize of

Avalanche

Observed Size of Slab Avalanche

0 2 3

Excl. SN38 Excl. SN38 Excl. SN38 Incl. SN38 Excl. SN38 Incl. SN38

0 119 119 6 5 4 4

2 2 2 2 3 0 0

3 0 0 0 0 0 0

Total 121 121 8 8 4 4

ProportionCorrect

119/121 119/121 2/8 3 0/4 0/4

PercentCorrect

98% 98% 25% 38% 0% 0%

relationship to MxN (Figure 9.1) indicating the importance of non-monotonic relationships

between avalanche forecasting variables.

9.3.3 Cariboo and Monashee Mountains

During the three winters that persistent weak layers were monitored in the Cariboo

and Monashee Mountains, measured or interpolated values of SN38 were obtained for

weak layers that produced more than one slab avalanche between 14 February 1993 and

30 March 1993, between 30 December 1993 and 22 March 1994 and between 10 January

1995 and 18 February 1995 for a total of 168 days. The daily avalanche activity for the

weak layers that produced one or more dry slab avalanche is described in Sections 6.6 and

6.9.

For the Cariboo and Monashee Mountains, the variables selected based on their

correlations with MxN (Table 9.1) are avalanche activity over the previous two days,

MN2, barometric pressure, PB, maximum temperature, Tmax, 24 h precipitation, HNW

and storm precipitation, HSTW. Wind speed, WS, and wind direction, WD, and height of

snowpack, HS, are selected from the box plots in Figure 9.1. The natural stability index,

SN38, is included to determine if it has predictive value in combination with the other

variables selected from Table 9.1 and Figure 9.1. Eliminating the days for which one or

more of these eight predictor variables is missing reduces the data set to 94 days.

Unfortunately, these data are highly unbalanced since there are only seven days with

persistent avalanches. Results from the classification tree models are summarized in

Table 9.4.

Using the eight forecasting variables MN2, WS, Tmax, WD, HNW, HSTW, HS and

PB, the classification tree algorithm selects MN2, WS, Tmax and WD as predictors for the

94 days mentioned previously (Model N-C-9E). The model achieves a misclassification

rate of 7/94 by classifying all days as non-avalanche days. Including SN38 with the other

eight variables results in Model N-C-9I which also classifies all avalanche days as non-

avalanche days. The data set is simply too small and too unbalanced to give interesting

results.

208

Table 9.4 Classification Trees for Daily Maximum Size of Natural Avalanches ofPersistent Slabs in the Cariboos and Monashees

ModelName Forecasting Variables1

No. ofTerminalNodes

Residual MeanDeviance

Mis-classification

rate

N-C-9E MN2, WS, Tmax, WD,HNW, HSTW, HS, PB

6 0.37 7/94

N-C-9I MN2, WS, SN38, Tmax,HNW, HSTW, HS, WD, PB

6 0.37 7/94

N-C-7 MN2, Tmax, HS, HSTW,HNW, PB, SN38

8 0.27 9/150

1 Variables marked in bold are selected by the classification tree algorithm from thoselisted and used to build the tree.

However, the size of the data set can be increased by eliminating the variables WS

and WD which are missing (due to riming problems with the anemometer and wind vane)

on more days than the other variables, except for SN38 which must be included to be

assessed (Table 9.1). Eliminating WS and WD increases the data set to 150 days including

13 avalanche days. Using the seven forecasting variables MN2, Tmax, HNW, HSTW, HS,

PB and SN38, the classification tree algorithm selects MN2, Tmax, HS, HSTW and rejects

SN38 along with HNW and PB for Model N-C-7 (Figure 9.4). Hence, for this larger data

set, SN38 does not contribute to a reduced misclassification rate for avalanche activity. As

shown in Table 9.5, all of the 137 days without avalanches and all of the four days with

class 3 or larger avalanches are correctly classified. However, all of the nine days with

class 1 or 2 avalanches are misclassified.

In summary, SN38 shows predictive value for natural avalanche activity in the Purcell

Mountains but not in the Cariboo and Monashee Mountains. This inconsistent result is

likely a consequence of the number of reports of natural avalanches of persistent slabs

being limited due to difficulty approaching natural avalanches—many of which start in

very steep terrain—and correctly identifying the failure plane. However, this difficulty

does not apply to skier-triggered avalanches considered in the next section.

209

210

Table 9.5 Contingency Table for Daily Maximum Size of Natural Avalanches ofPersistent Slabs in the Cariboo and Monashee Mountains

Modelled Size ofAvalanche

Observed Size of Slab Avalanche

0 1 2 ≥3

0 137 2 5 0

1 0 0 0 0

2 0 0 0 0

>2 0 0 2 4

Total 137 2 7 4

ProportionCorrect

137/137 0/2 0/7 4/4

Percent Correct 100% 0% 0% 100%

Figure 9.4 Classification tree for the daily maximum size of natural avalanches ofpersistent slabs in the Cariboo and Monashee Mountains based on 150 days from thewinters of 1992-93 to 1994-95. For each split, the left branch denotes days for which the“less than” criterion is true.

9.4 Forecasting Variables for Skier-Triggered Avalanches of Persistent

Slabs

This section assesses relationships between the daily maximum size of

skier-triggered persistent dry slab avalanches, MxS, and

meteorological variables PB, Ta, Tmin, Tmax, WS, WD, HN or HNW, HST or

HSTW, HS and PF which are available to the forecasters and ski guides on the

morning of the day,

the extrapolated skier stability index SK38 for persistent slabs,

MS1 which is the size class of the largest skier-triggered persistent slab on the

previous day, and

MS2 which is the sum of the size classes of the largest skier-triggered persistent

slabs from the two previous days.

As with the previous section for natural avalanches, variables are selected based on

either monotonic relationships detected with correlations or with non-monotonic

relationships apparent in box plots. The selected variables are used in multivariate

classification trees to assess the predictive value of SK38 for forecasting skier-triggered

persistent slabs.

Rank correlation coefficients, R, are presented in Table 9.6 between MxS and the

variables listed above based on data from the winters of 1992-93, 1993-94 and 1994-95.

The meteorological variables are the same as those used for the correlations with natural

avalanches in Section 9.2.

In the Purcell Mountains, MS1, MS2, BP, HN, HST and SK38 correlate significantly

(p < 0.05) with MxS. The negative correlation of barometric pressure and SK38 with MxS

implies that skier-triggering of persistent slabs tends to increase as barometric pressure

and SK38 decrease. The positive correlations of HN and HST with MxS imply that

skier-triggering of persistent slabs tends to increase as 24 h and storm snowfall increase.

In the Cariboo and Monashee Mountains, previous skier-triggered avalanche activity

(MS1 and MS2) as well as air temperature (Ta, Tmin and Tmax) and SK38 show significant

211

correlations with MxS (p < 0.05). Air temperature (Ta, Tmin and Tmax) and SK38 are

negatively and significantly correlated with MxS. Tmax has a stronger correlation with

MxS than Ta or Tmin and is used in subsequent analysis. The negative correlation may be

due to factors such as an increase in skier-triggered persistent slabs during the clearing and

cooling after a storm when skiing in avalanche terrain often resumes or increases, or a

reduction in skier-triggered persistent slabs in late winter and spring when air temperatures

rise and persistent weak layers are less common.

Box graphs are presented in Figure 9.5 to assess non-monotonic relationships of the

forecasting variables with MxS. MS2 is shown in the box plots although MS1 would

212

Table 9.6 Spearman Rank Correlations Between Forecasting Variables and the DailyMaximum Size of a Skier-Triggered Avalanche Involving a Persistent Slab

Forecasting Variables1Purcell Mountains Cariboo and

Monashee Mtns.

N R p N R p

Max. Skier-Trig. Slab on Previous Day(MS1)

294 0.14 0.02 355 0.50 <10-6

Sum of Max. Skier-Trig. Slab onPrevious 2 Days (MS2)

293 0.18 10-3 354 0.54 <10-6

Barometric Pressure (PB) 261 -0.12 0.05 353 -0.08 0.13

Air Temperature (Ta) 268 -10-3 0.98 327 -0.13 0.02

24 hr Min. Air Temperature (Tmin) 267 0.01 0.93 332 -0.12 0.03

24 h Max. Air Temperature (Tmax) 267 -0.07 0.24 334 -0.15 0.01

Wind Speed (WS) 264 0.01 0.82 258 0.05 0.44

Wind Direction (WD) 75 0.09 0.44 258 0.01 0.85

Height of 24 h Snow (HN) 264 0.15 0.01 - - -

Height of Storm Snow (HST) 255 0.15 0.01 - - -

Water Equiv. of 24 h Precip. (HNW) - - - 320 0.03 0.52

Water Equiv. of Storm Precip. (HSTW) - - - 320 0.04 0.44

Height of Snowpack (HS) 268 0.08 0.17 356 -0.06 0.23

Foot Penetration (PF) 263 0.08 0.20 - - -

Skier Stability Index (SK38) 138 -0.29 <10-3 168 -0.49 <10-6

1 Variables are measured as noted in Table 9.1.

probably have worked as well since MS1 and MS2 show comparable correlation

coefficients in Table 9.6. Ta and Tmin are excluded since they show similar but weaker

correlations than Tmax with MxS. Foot penetration is also excluded since only a

monotonic relationship is likely and the correlation with MxS is not significant (Table 9.6).

213

Figure 9.5 Box plots of the daily maximum size of a skier-triggered persistent slab againstvarious forecasting variables showing median (small rectangle), lower and upper quartiles(box) and minima and maxima (whiskers). Boxes for the Purcell Mountains near BobbyBurns Lodge are unshaded and boxes for the Cariboos and Monashee Mountains nearBlue River, BC are shaded. Precipitation values are height of snowfall in cm for thePurcell Mountains and height of melted precipitation in mm for the Cariboo and MonasheeMountains, 1992-93 to 1994-95.

HN and HST (cm) from the Purcell Mountains are plotted on the same graphs as HNW

and HSTW (mm) respectively from the Cariboo and Monashee Mountains.

In the Purcell Mountains, neither wind speed nor wind direction observed at Bobby

Burns Lodge (1370 m) shows a relationship with MxS in Figure 9.5. Reduced barometric

pressure, PB, is apparent on days with MxS = 1 compared to days with MxS = 0 and

MxS > 1. Height of snowpack, HS, shows a possible relationship with MxS. Hence, PB

and HS are included with Tmax, HN, HST and SK38 in the multivariate forecasting model

for the Purcell Mountains. The graph of SK38 in Figure 9.5 shows that most persistent

slabs were skier-triggered in the Purcell Mountains when SK38 is less than 1.5, which is the

critical value determined for SK and SK38 in Sections 6.8 and 6.9 respectively.

In the Cariboo and Monashee Mountains, neither the wind speed nor the wind

direction show a relationship with MxS in Figure 9.5. Maximum temperature, Tmax, and

barometric pressure, PB and SK38 were selected based on their correlations with MxS in

Table 9.6. This leaves 24 h precipitation, HNW, and storm precipitation, HSTW, and

height of snowpack, HS, all of which show increased median values for MxS = 2 than for

lower values of MxS. Consequently, Tmax, PB, SK38, HNW, HSTW and HS are used in the

multivariate forecasting model for the Cariboo and Monashee Mountains in the next

section. The graph of SK38 in Figure 9.5 shows that most persistent slabs were

skier-triggered in the Cariboo and Monashee Mountains when SK38 < 0.75. This is below

the critical value of 1.5 which was generally critical for most weak layers probably because

of the very low values of SK38 for the surface hoar layer buried on 10 February 1993 at the

Mt. St. Anne Study Plot (Section 6.9.2).

214

9.5 A Multivariate Forecasting Model for Skier-Triggered Avalanches

Involving Persistent Slabs

9.5.1 Cariboo and Monashee Mountains

In the Cariboo and Monashee Mountains there were 150 days with no missing

values for SK38, MS2, HST, BP, HS, HNW or Tmax. Excluding SK38 but including MS2,

BP, Tmax, HNW, HSTW and HS, the classification tree algorithm builds model S-C-E

(Table 9.7) which misclassifies MxS on 13 of the 150 days and has a residual mean

deviance of 0.37. Including SK38 results in the model S-C-I which also misclassifies 13

days but reduces the residual mean deviance to 0.33. The tree for this model is shown in

Figure 9.6.

The contingency tables for the tree without SK38 and the tree with SK38 are shown in

Table 9.8. The tree that includes SK38 misclassifies one less day with a class 1

skier-triggered slab than the tree that excludes SK38.

As was done for natural avalanches in the Purcell Mountains, the forecasting

variables can be ranked by simplifying the model and noting which variables are retained

as predictors. By increasing the cost-complexity factor, k, from 1 to 17, the variables are

ranked, in order of decreasing predictive value, MS2, SK38, HSTW, PB and HS

215

Figure 9.6 Classification tree for the daily maximum size of skier-triggered persistentslab in the Cariboo and Monashee Mountains based on data from the winters of1992-93 to 1994-95.

(Table 9.7). This approach ranks the non-monotonic relationships of HSTW and HS with

MxS (Figure 9.5) higher than the monotonic relationship of Tmax with MxS (Table 9.6),

indicating the importance of non-monotonic relationships in avalanche forecasting.

9.5.2 Purcell Mountains

The effect of SK38 on the misclassification rate for skier-triggered persistent slabs in

the Purcell Mountains can be assessed by considering classification trees developed with

and without SK38 from the same set of days. The selection of variables for the models is

based on correlations in Table 9.6 and box graphs in Figure 9.5. In Table 9.6, MS1, MS2,

PB, HN, HST and SK38 were significantly correlated with MxS (p < 0.05). In Figure 9.5,

216

Table 9.7 Classification Trees for Daily Maximum Size of Skier-Triggered PersistentSlabs in the Cariboos and Monashees, 1992-93 to 1994-95.

ModelName Forecasting Variables1

Cost-Complexity

Factor

No. ofTerminalNodes

ResidualMean

Deviance

Mis-classificatio

n rate

S-C-E MS2, HS, HSTW,Tmax, PB, HNW,

0 9 0.37 13/150

S-C-I MS2, SK38, HSTW, PB,HS, HNW, Tmax

0 9 0.33 13/150

S-C-1, 2, 3 MS2, SK38, HSTW, PB,HS, HNW, Tmax

1, 2, 3 7 0.36 15/150

S-C-4 MS2, SK38, HSTW, PB,HS, HNW, Tmax

4 6 0.39 15/150

S-C-5, 6, 7 MS2, SK38, HSTW, PB,HS, HNW, Tmax

5, 6, 7 5 0.44 15/150

S-C-8 MS2, SK38, HSTW, PB,HS, HNW, Tmax

8 4 0.49 20/150

S-C-9...16 MS2, SK38, HSTW, PB,HS, HNW, Tmax

9-16 3 0.60 20/150

S-C-17 MS2, SK38, HSTW, PB,HS, HNW, Tmax

17 2 0.73 29/150

1 Variables marked in bold are selected by the recursive partitioning algorithm for themodel.

Tmax appears to be of predictive value since it is generally lower on days when MxS is 1

than when MxS is 0. HS is also included since it appears greater when MxS is 1 than when

MxS is 0 or 2. MS2 is included in preference to MS1 since it exhibits a stronger correlation

in Table 9.4. Excluding the days in which any one of these variables is missing results in a

set of 133 days, including 16 days with skier-triggered persistent slabs.

From the variables, MS2, PB, HS, HST, Tmax and HN but excluding SK38, the

classification tree algorithm selects PB, HS, HST and Tmax but not HN as predictors of

MxS for Model S-P-E (Table 9.9). This model has 11 terminal nodes and a residual mean

deviance of 0.45. Unlike previous trees that excluded shear frame stability indices, this tree

(Figure 9.7) correctly classifies some avalanche days. As shown in the contingency table

(Table 9.9), this tree correctly classifies 110 of 117 non-avalanche days and 8 of 16

avalanche days for a misclassification rate of 15/133. Removing the subtrees that do not

reduce the misclassification rate (circled in Figure 9.7) increases the residual mean

deviance to 0.59.

217

Table 9.8 Contingency Table for Daily Maximum Size of Skier-Triggered PersistentSlabs in Cariboo and Monashee Mountains, 1992-93 to 1994-95.

PredictedSize of

Avalanche

Observed Size of Slab Avalanche

0 1 2

Excl. SK381 Incl. SK38

2 Excl. SK381 Incl. SK38

2 Excl. SK381 Incl. SK38

2

0 114 113 6 7 1 1

1 0 2 15 16 1 1

2 3 2 2 0 8 8

Total 117 117 23 23 10 10

ProportionCorrect

114/117 113/117 15/23 16/23 8/10 8/10

PercentCorrect

97% 97% 65% 70% 80% 80%

1 Predictions based on Model S-C-E which excludes SK38 as a forecasting variable.2 Predictions based on Model S-C-I which includes SK38 as a forecasting variable.

218

Figure 9.7 Classification tree for the daily maximum size of skier-triggered persistentslabs in the Purcell Mountains using meteorological forecasting variables but excludingSK38. Data are from the winters of 1992-93 to 1994-95.

Table 9.9 Classification Trees Results for Daily Maximum Size of Skier-TriggeredPersistent Slabs in the Purcell Mountains, 1992-93 to 1994-95.

ModelName Forecasting Variables

Cost-Complexity

Factor

No. ofTerminalNodes

ResidualMean

Deviance

Mis-classification

rate

S-P-E PB, HS, HST, Tmax,HN, MS2

0 11 0.45 15/133

S-P-I SK38, PB, HST, Tmax,HS, MS2, HN

0 12 0.42 12/133

S-P-1 SK38, PB, HST, Tmax,HS, MS2, HN

1 10 0.44 12/133

S-P-2 SK38, PB, HST, Tmax,HS, MS2, HN

2 9 0.45 12/133

S-P-3 SK38, PB, HST, Tmax,HS, MS2, HN

3 7 0.49 12/133

S-P-4, 5 SK38, PB, HST, Tmax,HS, MS2, HN

4, 5 5 0.57 12/133

S-P-6, 7,8, 9

SK38, PB, HST, Tmax,HS, MS2, HN

6, 7, 8, 9 4 0.64 14/133

S-P-10 SK38, PB, HST, Tmax,HS, MS2, HN

10 2 0.80 16/133

1 Variables marked in bold are selected by the classification tree algorithm from thoselisted for the model.

Including SK38 with the variables MS2, PB, HS, HST, Tmax and HN yields Model

S-P-I which reduces the residual mean deviance from 0.45 to 0.42 and improves the

misclassification rate from 15/133 to 12/133. This model correctly classifies 112 of 117

non-avalanche days and 9 of 16 avalanche days. Removing the subtrees that do not reduce

the misclassification rate (circled in Figure 9.8) increases the deviance to 0.57.

Although including SK38 (Model S-P-I) improves the overall misclassification rate, it

misclassifies 4 of 9 days with MxS = 2 compared to the model (M-P-E) without SK38

which misclassifies only 1 of 9 days with MxS = 2 (Table 9.10). While it is more important

for backcountry skiing operations to correctly predict days with class 2 avalanches than

days with no avalanches or class 1 avalanches, the classification tree algorithm weights all

values of MxS equally.

The variables in Model S-P-I can be ranked by increasing the cost-complexity factor

from 1 to 10 thereby simplifying the trees and retaining the most important variables at

each step (Table 9.9). Using this procedure, the variables in order of decreasing predictive

value are SK38, PB, HST, Tmax, HS, MS2, HN and MS2. Notably, Tmax, which did not

correlate significantly with MxS, ranked higher than HN or MS2 which did. This highlights

Figure 9.8 Classification tree for the daily maximum size of skier-triggered persistent slabs in the Purcell Mountains using meteorological forecasting variablesand including SK38. Data are from the winters of 1992-93 to 1994-95.

219

the relevance of choosing a forecasting model such as nearest neighbours or classification

trees which can include non-monotonic relationships.

Table 9.10 Contingency Table for Daily Maximum Size of Skier-TriggeredPersistent Slab in Purcell Mountains, 1992-93 to 1994-95.

PredictedSize of

Avalanche

Observed Size of Slab Avalanche

0 1 2 >2

Excl.SK38

1Incl.SK38

2Excl.SK38

1Incl.SK38

2Excl.SK38

1Incl.SK38

2Excl.SK38

1Incl.SK38

2

0 110 112 5 1 1 4 2 2

1 0 2 0 4 0 0 0 0

2 7 3 0 0 8 5 0 0

>2 0 0 0 0 0 0 0 0

Total 117 117 5 5 9 9 2 2

ProportionCorrect

110/117 112/117 0/5 4/5 8/9 5/9 0/2 0/2

PercentCorrect

94% 96% 0% 80% 89% 56% 0% 0%

1 Predictions based on Model S-P-E which excludes SK38 as a forecasting variable.2 Predictions based on Model S-P-I which includes SK38 as a forecasting variable.

9.6 Summary

Rank correlations, box graphs and multivariate classification trees were used to

assess the merit of shear frame stability indices, SN38 and SK38, for forecasting slab

avalanches involving persistent slabs.

For natural avalanches in the Purcell Mountains, SN38 showed promise based on the

correlations and box plots, whereas in the Cariboo and Monashee Mountains, SN38 showed

no predictive value. However, these analyses are not conclusive since the data for natural

avalanches of persistent slabs are highly unbalanced. Either there are few natural

avalanches with persistent failure planes, or difficulty with identifying the failure plane of

natural avalanches resulted in too few reports of natural avalanches with persistent failure

planes.

220

Fortunately, skier-triggered slabs are not observed from a distance and the reporting

of the failure plane is more rigorous. In both the Purcells and in the Cariboo and

Monashee Mountains, correlations and box plots indicate that SK38 is a useful predictor of

skier-triggered persistent slabs. Further, including SK38 as a forecasting variable improved

the number of days that multivariate classification trees correctly classified the size of the

largest skier-triggered persistent slab in both areas. Also, the classification tree algorithm

ranked SK38 as the most or the second most important forecasting variable in both areas.

However, the selection of variables was limited to SK38 and meteorological variables

from fixed sites that are available most mornings, as well as an index of previous avalanche

activity. Although avalanche workers report the results of intermittent snowpack

observations and tests from varying locations, such results were excluded from these

analyses which were limited to variables available daily from consistent locations. Expert

systems under development (e.g. Schweizer and Föhn, 1995) are likely to prove better

suited to including roving snowpack observations in a multivariate forecasting model.

The multivariate forecasting models presented in Sections 9.3 and 9.5 are of limited

value for operational forecasting since they are based on only 94 to 150 days. This is in

contrast to a nearest neighbour model that is now used operationally in Switzerland

(Buser, 1989) which is based on 20 years of data, and the nearest neighbour model being

tested at Kootenay Pass in BC (McClung and Tweedy, 1994) which is based on 10 years

of data. Backcountry skiing operations will require many years of data before such models

are operational practical, particularly since predictive accuracy is needed for large

skier-triggered avalanches which are infrequent.

Also, data-based models should be assessed with different data than those used to

build the model (e.g. Blattenberger and Fowles, 1995a). The data sets for the various

forecasting trees in this chapter were too small to withhold a portion of the data for such

an independent assessment.

However, the data sets for skier-triggered slabs were sufficient to determine that

multivariate forecasting models based on previous avalanche activity and common

meteorological variables can be improved by including SK38 in the model.

221

10 CONCLUSIONS

10.1 Field and Finite Element Studies of the Shear Frame Test

Most shear strengths from shear frame tests can be assumed to be normally distributed

since only 4 to 8 of 28 sets of 30 or more tests showed evidence of non-normality

(Section 4.2).

Coefficients of variation for shear frame tests averaged 0.15 and 0.18 from level study

plots and avalanche start zones respectively (Section 4.3). These values are less than

the 0.25 reported in previous studies and thus reduce the number of tests required to

achieve a specified level of precision. To achieve 10% precision at the 0.10

significance level, 8, 11 and 19 tests are required for shear frame data with coefficients

of variation of 0.15, 0.18 and 0.25, respectively.

In addition to the stress concentrations associated with the shear frame's rear

cross-member and intermediate fins, cutting along the front and back of the frame

with a blade notches the weak layer, thereby causing substantial stress concentrations

(Section 5.3). However, such cutting is essential to ensure that a specimen of known

size is tested.

Finite element studies showed that placing the frame a few mm above the weak layer

reduces stress concentrations compared to placing the bottom edges of the frame in

the weak layer (Section 5.4). This is consistent with field studies that showed strength

increases of 10-20% when the frame was placed 2-5 mm above the weak layer

compared to tests with the lower edges of the frame placed in the weak layer

(Section 4.8). Although frame placements 2-5 mm above the weak layer are

recommended, under certain snowpack conditions frames must be placed in the weak

layer or more than 5 mm above the weak layer to obtain planar failures in the weak

layer being tested.

For shear frame tests in which the shear frame is placed a few mm above the weak

layer, stiffer snow above the weak layer tends to reduce stress concentrations

(Section 5.5).

222

Shear frame tests that resulted in divots more than 10 mm deep under the rear

compartment of the frame yielded strengths significantly greater than tests with planar

fractures (Section 4.4). No significant effect could be detected for 10 other common

shapes of non-planar fracture surfaces.

With consistent loading rates and frame placement technique, there was no significant

difference in mean strengths obtained by different experienced shear frame operators

(Section 4.9).

Faster loading rates tended to reduce the strengths from shear frame tests. However,

the effect of loading rate on strength diminished for mean strengths less than 1 kPa

and for loading times less than 1 second (Section 4.5).

The first two tests in a set of shear frame tests were significantly more variable than

subsequent tests (Section 4.6). Rejecting the first two tests will therefore reduce

variability.

Delays of up to 3 minutes between placing the frame and pulling the frame did not

affect the resulting shear strengths (Section 4.7).

Shear frames with larger areas resulted in lower mean strengths than smaller frames

(Section 4.10), as shown in previous studies. Although strengths obtained with larger

frames usually showed reduced variability compared to smaller frames, the reduction

was not statistically significant for frames with areas of 0.01, 0.025 and 0.05 m2.

Increasing the number of cross-members while keeping the overall dimensions of the

frame constant increased the number of stress concentrations (Section 5.6) and

reduced the mean shear strength (Section 4.12.1).

Compared to the 0.025 m2 shear frame with three active cross-members used as a

standard in the present study, the Swiss shear frame is constructed of thicker metal

and consequently is heavier. It resulted in increased shear strengths compared to the

standard frame (Section 4.12.3).

Compared to the standard frame, the finger-fin shear frame resulted in decreased shear

strengths due to reduced stress concentrations, but operators had difficulty placing the

223

finger-fin frame a certain distance above the weak layer, a practice that is commonly

required to obtain planar fractures in certain weak layers (Section 4.12.4).

10.2 Shear Strength of Weak Layers

Regressions for estimating the shear strength of common microstructures from density

are presented in Section 6.2. For those weak layers that are too thin for density

measurements, Section 6.3 provides a graph for estimating mean strength for common

microstructures from classes of hand hardness.

Persistent weak planes consisting of surface hoar or facets showed less strength

increase with increased normal load than reported previously (Section 4.11). Since the

normal load effect for persistent weak layers was not significant, it was taken to be

negligible. This may result in conservative stability indices (lower than otherwise) for

thick, dense slabs.

10.3 Shear Frame Stability Indices

Values of shear frame stability index for natural avalanches SN, which differ from S'

developed by Föhn (1987a) only in the normal load adjustment, are presented for

various slopes that avalanched naturally and those that did not avalanche

(Section 6.5). For each of four slopes that avalanched with high values of SN, warming

or ambient temperatures near 0oC are likely explanations, indicating that SN cannot

predict avalanches under such conditions.

SN38 is obtained by calculating SN for a 38° inclination typical of start zones. Most

natural avalanches occurred on surrounding slopes when SN38 was less than 2.8.

However, non-avalanche days were common for a wide range of values of SN38

(Section 6.6). Based on univariate and multivariate analyses of data from three winters

with a limited number of natural avalanches of persistent slabs, SN38 showed promise

for forecasting natural avalanches in the Purcell Mountains. A similar relationship

between SN38 and natural avalanche activity was not detected in the Cariboo and

224

Monashee Mountains (Sections 9.2 and 9.3). However, in both forecast areas,

identifying the failure plane and occurrence date of natural avalanches was difficult

since many natural avalanches were observed from a distance.

The transitional stability of SN and SN38 falls well above 1 suggesting that a critical

stress failure criterion is not well suited to predicting natural avalanching.

An empirical formula for estimating ski penetration from slab density and thickness

was incorporated into a formula derived by Föhn (1987a) resulting in a stability index

for skier-triggering, SK, which has a reduced number of false stable predictions for

skier-tested avalanche slopes (Section 6.8).

SK38 is obtained by calculating SK for a 38° inclination typical of start zones. In

Section 6.9, SK38 is shown to be better predictor of skier-triggered avalanches in

surrounding terrain than is SN38 for natural avalanches. Also, values of SK and SK38

between 1 and 1.5 correspond to transitional stability for both test slopes and

surrounding terrain, indicating that the critical stress failure criterion upon which SK

and SK38 are based is effective for skier-triggered avalanches.

Univariate and multivariate analyses for three winters at both forecast areas showed

SK38 to be a better predictor for skier-triggered slabs than common meteorological

observations (Sections 9.4 and 9.5). Including SK38 as a forecasting variable improved

the number of days that multivariate forecasting models correctly classified the size of

the largest skier-triggered persistent slab in both areas. However, the selection of

variables excluded roving snowpack observations and tests such as profiles, shovel

tests, compression tests, and rutschblock tests normally done only when deemed

necessary.

Differences in the initial size of surface hoar crystals between two study sites affect

stability. If surface hoar crystals at a particular site are substantially larger than at a

second site, then stability will tend to be lower at the first site and remain that way for

a period of weeks.

225

10.4 Rutschblock Results

Closely spaced rutschblock tests on nine avalanche slopes illustrate snowpack and

terrain factors that affect rutschblock scores (Section 7.2). Sites near the top of

slopes, near trees, over rocks and at pillows of wind-deposited snow sometimes

exhibited rutschblock scores and/or failure planes quite different from the remainder

of the slope. Even when avoiding such sites, rutschblock scores two steps above the

slope median occurred occasionally, indicating the merit of using other sources of

information such as profiles to confirm or question the results of one or two

rutschblock tests.

The percentage of skier-triggered persistent slabs on skier-tested avalanche slopes

decreased from over 80% to 33% as median rutschblock scores increased from 2 to 5

(Section 7.3). Three of nine skier-tested slopes with median scores of 7 for persistent

slabs were skier-triggered, indicating that the median rutschblock score is, by itself,

not a completely reliable indicator of stability. For non-persistent weak layers, no slabs

were skier-triggered on 10 slopes with median rutschblock scores above 3.

The frequency of avalanching for non-persistent slabs with rutschblock scores of 2, 3

and 4 was approximately the same as for persistent slabs with rutschblock scores of 4,

5 and 6, respectively. Consequently, the interpretation of rutschblock scores should

depend on whether the weak layer is persistent or not.

An empirical relationship between skier-stability index, SK, and median rutschblock

scores was determined (Section 7.4). However, this relationship does not apply on

slopes of less than 20° where SK usually predicts higher stability than the rutschblock

test. Since SK is based on shear failure, it is argued that the primary fractures

sometimes initiated by skiers on slopes of less than 20° are compressive.

Rutschblocks on safe study slopes are shown to have predictive value for

skier-triggering of particular persistent layers on surrounding slopes (Section 7.6).

However, as for rutschblock tests on avalanche slopes, the particular persistent weak

226

layer is sometimes skier-triggered when the rutschblock score for that layer on the

study slope indicates stability.

10.5 False Stable Predictions

Case studies illustrated that stability tests where snowpack conditions are judged

typical of start zone are occasionally misleading and cannot predict avalanches

triggered at localized weaknesses or remotely from sites with a less stable snowpack

than the start zone (Chapter 8). These case studies confirm the advice of other authors

that the results of such tests should be interpreted together with knowledge of terrain,

snow distribution and mesoscale stability trends based on regular weather, study site

and avalanche observations, rather than on a stand-alone basis.

227

11 RECOMMENDATIONS FOR FURTHER RESEARCH

The relationship of observable microstructural properties of buried surface hoar

layers such as mean grain size and change in mean grain size to the strength of these layers

should be studied through a combination of microphotography and strength tests. The

changes of particular layers over time on slopes and at level sites would be helpful since

creep on steeper slopes may increase the number of bonds per crystal over time by

inclining the crystals. An increase in the number of bonds per crystal and consequently an

increase in strength on steeper slopes may explain why avalanches are sometimes triggered

remotely from level areas or shallow slopes, sometimes after the steeper slopes have

apparently stabilized. The strength tests in the level areas should not be restricted to shear

tests since the primary fractures at such sites probably involve compression.

Based on shear frame tests in well chosen study sites, the skier-stability index, SK38,

is an effective predictor of skier-triggered persistent slabs in surrounding terrain.

However, because of the costs associated with skilled avalanche technicians and

transportation to study sites to do shear frame tests, there are economic advantages to

reducing the frequency of the tests. Although interpolating SK38 between test days proved

useful for assessing the merit of SK38 for predicting past skier-triggered avalanche activity,

operational use of SK38 will require either that shear frame tests in study sites be

conducted frequently, perhaps every third day, which is expensive, or that SK38 be

estimated based on last measured value and on easily measured field parameters since the

last test day. Since SK38 is based on shear strength of the persistent weak layer, load (slab

weight per unit area) and slab thickness based on snowfall and settlement, predicting SK38

will require estimates of

1. changes in shear strength of the persistent weak layer based on easily measured

parameters such as temperature of the weak layer, temperature gradient across the

weak layer, load and microstructure,

2. increases in load based on daily measurements of snowfall or precipitation from a

easily accessible study site or automatic weather station, and

228

3. settlement based on snowfall, load, slab density, microstructure and, perhaps,

temperature (e.g. Navarre, 1975; Armstrong, 1980; Brun and others, 1989).

While the load and settlement can be estimated from easily measured field

parameters, predictive models for changes in shear strength of thin persistent weak layers

are needed. Such models will permit less frequent visits to study sites to test persistent

weak layers with the shear frame and may prove cost-effective for backcountry avalanche

forecasting operations.

Forecasting models based on data, knowledge or both should be developed for

backcountry forecasting programs. The skier-stability index, SK38, and/or rutschblock tests

at regular intervals on study slopes should be incorporated into such models. However,

data-based models will require additional years of systematic snowpack tests and weather

observations. Various snowpack tests such as rutschblock tests, compression tests, shovel

tests and profiles done at sites in and near avalanche start zones may prove useful for

expert systems particularly when coupled with past, present and forecast weather

(Schweizer and Föhn, 1995). Such models can potentially assist forecasters either by

possibly identifying an overlooked unstable condition, or by supporting the forecaster's

decisions in the event of an unexpected avalanche.

The fractures that release slab avalanches sometimes propagate from localized weak

areas near rocks, bushes, etc. Tests at such potential trigger points are unlikely to

correlate with occurrences of remote triggering tests since the snowpack properties at

such sites are highly variable. However, a propagation index based on the ratio of the

strain energy capacity of the slab to the fracture toughness of the weak layer (Jamieson

and Johnston, 1992b) based on study site measurements may prove practical upon further

investigation.

229

REFERENCES

Adams, E.E. and R.L. Brown. 1982. Further results on studies of temperature-gradient

metamorphism. Journal of Glaciology 28(98), 205-210.

Akitaya, E. 1974. Studies on depth hoar. Contributions from the Institute of Low

Temperature Science, Series A, 26, 1-67.

Akitaya, E., 1975. Studies on depth hoar, Snow Mechanics, Proceedings of the

Grindewald Symposium, International Association of Hydrological Sciences,

Washington, D.C., Publication No. 114, 42-48.

Armstrong, R.L. 1977. Continuous monitoring of metamorphic changes of internal snow

structure as a tool in avalanche studies. Journal of Glaciology 18(81), 325-334.

Armstrong, R.L. 1980. An analysis of compressive strain in adjacent temperature-gradient

and equi-temperature layers in a natural snow cover. Journal of Glaciology, 26(94),

283-289.

Armstrong, R.L. 1981. Some observations on snowcover temperature patterns.

Proceedings of the Avalanche Workshop in Vancouver, 3-5 November 1980. National

Research Council of Canada Technical Memorandum 133, 66-75.

Bader, H.-P., H.U. Gubler, and Salm, B. Distribution of stress and strain rates in

snowpacks. Numerical Methods in Geomechanics (Innsbruck 1988) Swoboda (ed.)

1989 Balkema, Rotterdam.

Bader, H.P., and B. Salm. 1990 On the mechanics of snow slab release. Cold Regions

Science and Technology, 17, 287-300.

Ballard, G.E.H. and E.D. Feldt, 1965. A theoretical consideration of the strength of snow,

Journal of Glaciology, 6(43), 159-170.

Barry, R.G. and R.J. Chorley. 1987. Atmosphere, Weather and Climate. Methuen, New

York. 460 pp.

Blattenberger, G. and R. Fowles. 1995a. The road closure decision in Little Cottonwood

Canyon. Proceedings of the 1994 International Snow Science Workshop in Snowbird,

Utah. ISSW '94, P.O. Box 49, Snowbird, Utah, USA, 537-547.

230

Blattenberger, G. and R. Fowles. 1995b. Road closure to mitigate avalanche danger: a

case study for Little Cottonwood Canyon. International Journal of Forecasting 11,

159-174.

Bovis, M.J. Statistical forecasting of snow avalanches, San Juan Mountains, southern

Colorado, USA. Journal of Glaciology 18(78), 87-99.

Boyne, H and K. Williams. 1993. Analysis of avalanche prediction from meteorological

data at Berthoud Pass, Colorado. Proceedings of the International Snow Science

Workshop in Breckenridge, Colorado, October 4-8, 1992. ISSW '92 Committee, c/o

Colorado Avalanche Information Centre, 10230 Smith Road, Denver, Colorado, 80239

USA, 229-235.

Bradley, C.C., 1966. The snow resistograph and slab avalanche investigations.

Proceedings of the International Symposium on Scientific Aspects of Snow and Ice

Avalanches, Davos, April 5-10, 1965, International Association of Scientific

Hydrology, Publication 69, 251-260.

Bradley, C.C. and D. Bowles. 1967. Strength-load ratio, an index of deep slab avalanche

conditions. Physics of Snow and Ice, (H. Oura, Ed.) 1, Part 2, Institute of Low

Temperature Science, Hokkaido University, Japan, 1243-1253.

Bradley, C.C., R.L. Brown and T.R. Williams, 1977a. On depth hoar and the strength of

snow. Journal of Glaciology 18(78), 145-147.

Bradley, C.C., R.L. Brown and T.R. Williams, 1977b. Gradient metamorphism, zonal

weakening of the snowpack and avalanche initiation, Journal of Glaciology, 19(81),

335-342.

Breiman, L., J.H. Freidman, R.A. Olshen and C.J. Stone. 1984. Classification and

Regression Trees. Wadsworth and Brooks/Cole Advanced Books and Software, Pacific

Grove, CA, 358 pp.

Brown, R.L., 1977. A fracture criterion for snow, Journal of Glaciology, 19(81), 111-121.

Brown, R.L., T.E. Lang, W.F. St. Lawrence and C.C. Bradley, 1973. A failure criterion

for snow, Journal of Geophysical Research, 78(23), 4950-4958.

231

Brun, E., P. David, M. Sudul and G. Brunot. 1992. A numerical model to simulate

snow-cover stratigraphy for operational avalanche forecasting. Journal of Glaciology

38(128), 13-22.

Brun, E., E. Martin, V. Simon, C. Gendre and C. Coleou. 1989. An energy and mass

model of snow cover suitable for operational avalanche forecasting. Journal of

Glaciology, 35(121), 333-342.

Brun, E. and L. Rey. Field study on snow mechanical properties with special regard to

liquid water content. Avalanche Formation, Movement and Effects, Edited by B. Salm

and H. Gubler, International Association of Hydrological Sciences, Publication No.

162, 183-193.

Bucher, E., 1948. Contribution to the theoretical foundations of avalanche defense

construction, Snow, Ice and Permafrost Research Establishment, Translation 18, 1956,

99 pp.

Buser, O. 1983. Avalanche forecast with the methods of nearest neighbours: an interactive

approach. Cold Regions Science and Technology, 8(2) 155-163.

Buser, O. 1989. Two years experience of operational avalanche forecasting using the

nearest neighbours method. Annals of Glaciology, 13, 31-34.

Buser, O., P. Föhn, W. Good, H. Gubler and B. Salm. 1985. Different methods for the

assessment of avalanche danger. Cold Regions Science and Technology, 10, 199-218.

Buser, O., M. Bütler and W. Good. 1987. Avalanche forecast by the nearest neighbour

method. Avalanche Formation, Movement and Effects, Edited by B. Salm and H.

Gubler, International Association of Hydrological Sciences, Publication 162, 557-569.

CAA. 1994. Course Manual for Avalanche Safety Courses. Canadian Avalanche

Association. P.O. Box 2759, Revelstoke, BC, Canada, 128 pp.

CAA. 1995. Observation Guidelines and Recording Standards for Weather, Snowpack

and Avalanches. Canadian Avalanche Association. P.O. Box 2759, Revelstoke, BC,

Canada, 98 pp.

Chambers, J.M. and T.J. Hastie. 1992. Statistical Models in S. Wadsworth and Brooks.

Pacific Grove, California, 600 pp.

232

Clifford, A.A. 1973. Multivariate Error Analysis, A Handbook of Error Propagation and

Calculation in Many-Parameter Systems. John Wiley and Sons, New York. 112 pp.

Colbeck, S.C. 1983. Theory of metamorphism of dry snow. Journal of Geophysical

Research 88(C9), 5475-5482.

Colbeck, S.C., 1987. A review of metamorphism and the classification of seasonal snow

cover crystals, Avalanche Formation, Movement and Effects, Edited by B. Salm and H.

Gubler, International Association of Hydrological Sciences, Publication 162, 3-33.

Colbeck, S; Akitaya, E; Armstrong, R; Gubler, H; Lafeuille, J; Lied, K; McClung, D; and

Morris, E. 1990. International Classification for Seasonal Snow on the Ground.

International Commission for Snow and Ice (IAHS), World Data Center A for

Glaciology, U. of Colorado, Boulder, CO, USA.

Colbeck, S.C. 1991. The layered character of snow covers. Reviews of Geophysics 29(1),

81-96.

Conway, H. and J. Abrahamson, 1984. Snow stability index, Journal of Glaciology,

30(106), 321-327.

Conway, H. and J. Abrahamson, 1988. Snow-slope stability - A probabilistic approach,

Journal of Glaciology, 34(117), 170-177.

Curtis, J.O. and F.W. Smith. 1975. Stress analysis and failure prediction in avalanche

snowpacks. Snow Mechanics, Proceedings of the Grindewald Symposium,

International Association of Hydrological Sciences, Washington, D.C., Publication No.

114, 332-340.

Daniels, H.E. 1945. The statistical theory of the strength of bundles of threads.

Proceedings of the Royal Society of London, Series A, 183(995), 405-435.

Davis, R. K. Elder and D. Bouzaglou. 1993. Applications of classification tree

methodology to avalanche data management and forecasting. Proceedings of the

International Snow Science Workshop in Breckenridge, Colorado, October 4-8, 1992.

ISSW '92 Committee, c/o Colorado Avalanche Information Centre, 10230 Smith Road,

Denver, Colorado, 80239 USA, p. 126-133.

233

Davis, R. E. and K. Elder. 1994. Management and data analysis of weather and avalanche

records: Recent directions and perspectives with case studies. Proceedings of the 1994

Eastern Snow Conference, 143-150.

Davis, R. E. and K. Elder. 1995. Application of classification and regression trees:

selection of avalanche activity indices at Mammoth Mountain. Proceedings of the 1994

International Snow Science Workshop in Snowbird, Utah. ISSW '94, P.O. Box 49,

Snowbird, Utah, USA, 285-294.

de Montmollin. 1982. Shear tests on snow explained by fast metamorphism. Journal of

Glaciology 28(98), 187-198.

de Quervain, M., 1951. Strength Properties of a Snow Cover and Its Measurement, U.S.

Army Snow Ice and Permafrost Research Establishment, Translation 9, 9 pp.

de Quervain, M. 1958. On metamorphism and hardening of snow under constant pressure

and temperature gradient. International Association of Scientific Hydrology,

Proceedings of the General Assembly in Toronto, 1957, Vol. 4, 225-239.

de Quervain, M., 1963. On the metamorphism of snow. Ice and Snow: Properties,

Processes and Applications, M.I.T. Press, Cambridge, Mass., 377-390.

DenHartrog, S.L. 1982. Firn quake. Cold Regions Science and Technology 6, 173-74.

Fredston, J. and D. Fesler. 1994. Snow Sense: A Guide to Evaluating Snow Avalanche

Hazard. Alaska Mountain Safety Center, Inc., Anchorage, Alaska, 116 pp.

Föhn, P.M.B, 1987a. The stability index and various triggering mechanisms, Avalanche

Formation, Movement and Effects, Edited by B. Salm and H. Gubler, International

Association of Hydrological Sciences, Publication No. 162, 195-211.

Föhn, P.M.B. 1987b. The rutschblock as a practical tool for slope stability evaluation.

Avalanche Formation, Movement and Effects, IASH Publ. 162 (Symposium at Davos

1986), 223-228.

Föhn, P.M.B., 1989. Snowcover stability tests and the areal variability of snow strength,

Proceedings of the International Snow Science Workshop in Whistler, B.C., October

12-15, 1988, 262-273.

234

Föhn, P.M.B. 1993. Characteristics of weak snow layers or interfaces. Proceedings of the

International Snow Science Workshop in Breckenridge, Colorado, October 4-8, 1992.

ISSW '92 Committee, c/o Colorado Avalanche Information Centre, 10230 Smith Road,

Denver, Colorado, 80239 USA, p. 171-175.

Fukuzawa, T. and H. Narita. 1993. An experimental study on the mechanical behaviour of

a depth hoar layer under shear stress: preliminary report. Proceedings of the

International Snow Science Workshop in Breckenridge, Colorado, October 4-8, 1992.

ISSW '92 Committee, c/o Colorado Avalanche Information Centre, 10230 Smith Road,

Denver, Colorado, 80239 USA, 160-170.

Giraud, G. MEPRA, An expert system for avalanche risk forecasting. Proceedings of the

International Snow Science Workshop in Breckenridge, Colorado, (Oct. 1992),

97-104.

Gray, J.M.N.T, L.W. Morland and S.C. Colbeck. 1995. The effect of change in the

thermal properties on the propagation of a periodic thermal wave: Application to a

snow buried rocky outcrop. Journal of Geophysical Research, 100(B8) 15,267-15,279.

Griffith, A.A., 1920. The Phenomena of Rupture and Flow in Solids, Philosophical

Transactions of the Royal Society, London, Series A, Vol. 221, 163-198.

Gubler, H., 1978. Determination of the mean number of bonds per snow grain and of the

dependence of the tensile strength of snow on stereological parameters, Journal of

Glaciology, 20(83), 329-340.

Gubler, H. 1982. Strength of bonds between ice grains after short contact times. Journal

of Glaciology, 28(100), 457-473.

Gubler, H. and H.P. Bader. 1989. A model of initial failure in slab avalanche release.

Annals of Glaciology 13, 90-95.

Hachikubo, A. T. Fukuzawa and E. Akitaya. 1995. Formation rate of surface hoar crystals

under various wind velocities. Proceedings of the International Snow Science

Workshop at Snowbird, International Snow Science Workshop 1994, P.O. Box 49,

Snowbird, Utah 84092, USA, 132-137.

235

Haefeli, R., 1939. Snow Mechanics with Reference to Soil Mechanics. Snow and its

Metamorphism, Translation 14, Snow, Ice and Permafrost Research Establishment,

58-218.

Haefeli, R., 1963. Stress transformations, tensile strengths, and rupture processes of the

snow cover, Ice and Snow: Properties, Processes and Applications, Edited by W.D.

Kingery, M.I.T. Press, Cambridge, Massachusetts, 560-575.

Haefeli, R. 1967. Some mechanical aspects on the formation of avalanches. Physics of

Snow and Ice, (H. Oura, Ed.) Vol.1, Part 2, Institute of Low Temperature Science,

Hokkaido University, Japan, 1199-1213.

Holtz, D.H. and Kovacs, W.D. An Introduction to Geotechnical Engineering. Prentice

Hall, New Jersey, 733 pp.

Jamieson, J.B. 1989. In situ tensile strength of snow in relation to slab avalanches. MSc

thesis, Dept. of Civil Engineering, University of Calgary, 142 pp.

Jamieson, J.B. and C.D. Johnston, 1990. In situ tensile tests of snowpack layers, Journal

of Glaciology 36(122), 102-106.

Jamieson, J.B. and C.D. Johnston, 1992a. Snowpack characteristics associated with

avalanche accidents. Canadian Geotechnical Journal 29, 862-866.

Jamieson, J.B. and C.D. Johnston, 1992b. A fracture-arrest model for unconfined dry slab

avalanches. Canadian Geotechnical Journal 29, 61-66.

Jamieson, J.B. and C.D. Johnston, 1993a. Shear frame stability parameters for large scale

avalanche forecasting. Annals of Glaciology 18, 268-273.

Jamieson, J.B. and C.D. Johnston, 1993b. Rutschblock precision, technique variations and

limitations. Journal of Glaciology 39(133), 666-674.

Jamieson, J.B. and C.D. Johnston, 1993c. Experience with rutschblocks. Proceedings of

the International Snow Science Workshop in Breckenridge, Colorado, (Oct. 1992),

150-159.

Jamieson, J.B. and C.D. Johnston. 1995a. Monitoring a shear frame stability index and

skier-triggered slab avalanches involving persistent snowpack weaknesses.

236

Proceedings of the International Snow Science Workshop at Snowbird, International

Snow Science Workshop 1994, P.O. Box 49, Snowbird, Utah 84092, USA, 14-21.

Jamieson, J.B. And C.D. Johnston. 1995b. Interpreting rutschblocks in avalanche start

zones. Avalanche News, 46, 2-4.

Judson, A. and B.J. Erickson. 1973. Predicting avalanche intensity from weather data: a

statistical analysis. USDA Forest Service Research Paper RM-112, 12 pp.

Keeler, C.M., 1969. Some physical properties of alpine snow, U.S. Army Cold Regions

Research and Engineering Laboratory, Technical Note, 67 pp.

Keeler, C.M. and W.F. Weeks, 1968. Investigations into the mechanical properties of

alpine snow-packs, Journal of Glaciology, 7(5), 253-271.

Kristensen, K. and C. Larsson. 1995. An avalanche forecast program based on a modified

nearest neighbour method. Proceedings of the International Snow Science Workshop at

Snowbird, International Snow Science Workshop 1994, P.O. Box 49, Snowbird, Utah

84092, USA, 22-30.

Kry, P.R., 1975. The relationship between the visco-elastic and structural properties of

fine grained snow, Journal of Glaciology, 14(72), 479-500.

LaChapelle, E.R. 1980. The fundamental processes in conventional avalanche forecasting.

Journal of Glaciology 26(94), 75-84.

Lang, R. M., B.R. Leo and R.L. Brown. 1985. Observations on the growth process and

strength characteristics of surface hoar. Proceedings of the International Snow Science

Workshop at Aspen, October 1984. ISSW Workshop Committee, C/O Mountain

Rescue-Aspen, Inc., P.O. Box 4446, Aspen, CO, 81612, USA, 188-195.

Lipson, C. and N.J. Sheth. 1973. Statistical Design and Analysis of Engineering

Experiments. McGraw-Hill, New York, 518 pp.

Male, D.H., 1980. The seasonal snowcover. Dynamics of Snow and Ice Masses, Edited by

S.C. Colbeck, Academic Press, New York, 305-395.

Martinelli Jr., M., 1971. Physical properties of alpine snow as related to weather and

avalanche conditions, U.S. Dept. of Agriculture, Forest Service Research Paper

RM-64, 30 pp.

237

Mattson, D.E. 1981. Statistics: Difficult Concepts, Understandable Explanations. Mosby,

St. Louis, 480 pp.

McClung, D.M., 1977. Direct simple shear tests on snow and their relation to slab

avalanche formation, Journal of Glaciology, 19(81), 101-109.

McClung, D.M., 1979. Shear fracture precipitated by strain softening as a mechanism of

dry slab avalanche release, Journal of Geophysical Research, 84(B7), 3519-3526.

McClung, D.M., 1981. Fracture mechanical models of dry slab avalanche release, Journal

of Geophysical Research, 86(B11), 10783-10790.

McClung, D.M., 1987. Mechanics of snow slab failure from a geotechnical perspective,

Proceedings from the International Symposium on Avalanche Formation, Movement

and Effects, Davos, International Association of Hydrological Sciences, Publication

No. 162, New York, 475-507.

McClung, D.M. 1995. Computer assistance in avalanche forecasting. Proceedings of the

1994 International Snow Science Workshop in Snowbird, Utah. ISSW '94, P.O. Box

49, Snowbird, Utah, USA, 310-313.

McClung, D.M. and P.A. Schaerer. 1993. The Avalanche Handbook. The Mountaineers,

Seattle, 271 pp.

McClung, D.M. and J. Tweedy. 1994. Numerical avalanche prediction: Kootenay Pass,

British Columbia, Canada. Journal of Glaciology, 40(135), 350-358.

McClung, D.M. 1995. Computer assistance in avalanche forecasting. Proceedings of the

1994 International Snow Science Workshop in Snowbird, Utah. ISSW '94, P.O. Box

49, Snowbird, Utah, USA, 310-312.

Mears, A.I. 1992. Snow-avalanche hazard analysis for land-use planning and engineering.

Colorado Geological Survey, Bulletin 49, 55 pp.

Mellor, M., 1975. A review of basic snow mechanics. Snow Mechanics Symposium,

Proceedings of the Grindewald Symposium, April, 1974, International Association of

Hydrological Sciences, Washington, D.C., Publication 114, 1975, 251-291.

Mellor, M. and J.H. Smith, 1966. Strength studies of snow. Cold Regions Research and

Engineering Laboratory, Research Report 168, 14 pp.

238

Moore, Mark. 1982. Temperature gradient weakening of snowpacks near rain crusts or

melt-freeze layers. Presented at the 1982 International Snow Science Workshop in

Bozeman, Montana. Unpublished.

Morrall, J.F. and W.M. Abdelwahab. 1992. Estimating traffic delays and the economic

cost of recurrent road closures on rural highways. Logistics and Transportation Review

29(2), 159-177.

Munter. W. 1973. Kliene Schnee- und lawinenkunde. Skiführer in "Alpinismus", Heft 1,

1973, Heering-Verlag, München, 32 pp.

Munter, W. 1991. Neue Lawinenkunde, Ein Leitfaden für die Praxis. Schweizer

Alpen-Club, Bern, Switzerland, 200 pp.

Narita, H., 1980. Mechanical behaviour and structure of snow under uniaxial tensile stress,

Journal of Glaciology, 26(94), 275-282.

Narita, H., 1983. An experimental study on the tensile fracture of snow, Contributions

from the Institute of Low Temperature Science, Contribution No. 2625, Institute of

Low Temperature Science, Hokkaido University, Sapporo, Japan, 1-37.

Narita, H., T. Fukuzawa, N. Maeno and W. Ma. 1992. Sample shear deformation of snow

containing a weak layer. Presented at the Symposium on Snow and Snow-Related

Problems in Nagaoka, Japan, 14-18 September, 1992. Unpublished.

Navarre, J.P. 1975. Model unidimensional d'evolution de la neige déposée. Modèle

perce-neige. Météorologie 4(3), 103-120.

NRCC and CAA, 1989. Guidelines for Weather, Snowpack and Avalanche Observations.

National Research Council of Canada and Canadian Avalanche Association, NRCC

Technical Memorandum 132, 53 pp.

Obled, C. and W. Good. 1980. Recent developments of avalanche forecasting by

discriminant analysis techniques: a methodological review and some applications to the

Parsenn Area (Davos, Switzerland). Journal of Glaciology 25(92), 315-346.

Palais, J.M. 1984. Snow stratigraphic investigations at Dome C Antartica: A study of

depositional and diagenetic processes, Report 1984-No. 78, 121 pp. Institute for Polar

Studies, Ohio State University, Columbus.

239

Palmer, A.C. and J.R. Rice, 1973. The growth of slip surfaces in the progressive failure of

overconsolidated slay, Proceedings of the Royal Society, Vol. A 332, 527-548.

Perla, R.I. 1970. On contributory factors in avalanche hazard evaluation. Canadian

Geotechnical Journal 7(4), 414-419.

Perla, R.I., 1975. Stress and fracture of snow slabs, Snow Mechanics Symposium,

Proceedings of the Grindewald Symposium, International Association of Hydrological

Sciences, Washington, D.C., Publication 114, 208-221.

Perla, R.I., 1977. Slab avalanche measurements, Canadian Geotechnical Journal, 14(2),

206-213.

Perla, R.I., 1980. Avalanche release, motion and impact. Dynamics of Snow and Ice

Masses, Edited by S.C. Colbeck, Academic Press, New York, 397-462.

Perla, R.I. and E.R. LaChapelle, 1970. A theory of snow slab failure, Journal of

Geophysical Research, 75(36), 7619-7627.

Perla, R.I. and M. Martinelli, Jr., 1976. Avalanche Handbook, U.S. Dept. of Agriculture,

Agriculture Handbook 489, 238 pp.

Perla, R.I., T.M.H. Beck and T.T. Cheng, 1982. The shear strength index of alpine snow,

Cold Regions Science and Technology, 6, 11-20.

Perla, R.I., and T.M.H. Beck, 1983. Experience with shear frames, Journal of Glaciology,

29(103), 485-491.

Perla, R.I. and C.S.L. Ommanney, 1985. Snow in strong or weak temperature gradients.

Part I: Experiments and qualitative observations, Cold Regions Science and

Technology, 11, 23-35.

Rice, J.R., 1973. The initiation and growth of shear bands, Proceedings of the Symposium

on the Role of Plasticity in Soil Mechanics, Cambridge University Engineering Dept.,

Cambridge, England, 263-277.

Roch, A., 1956. Mechanism of avalanche release, U.S. Army, Snow, Ice and Permafrost

Research Establishment, Hanover, New Hampshire, Translation 52, 11 pp.

240

Roch, A., 1966a. Les declenchements d'avalanche, Proceedings of the International

Symposium on Scientific Aspects of Snow and Ice Avalanches, Davos, April 1965,

182-183.

Roch, A., 1966b. Les variations de la resistance de la neige, Proceedings of the

International Symposium on Scientific Aspects of Snow and Ice Avalanches,

Gentbrugge, Belguim, International Association of Hydrological Sciences, 182-195.

Salm, B., 1971. On the rheological behaviour of snow under high stresses. Contributions

from the Institute of Low Temperature Science, Hokkaido University, Series A, No.

23, 43 pp.

Salm, B., 1981. Mechanical properties of snow. Proceedings of a Workshop on the

Properties of Snow, Snowbird, Utah, U.S. Army, Cold Regions Research and

Engineering Laboratory, Hanover, New Hampshire, Special Report 82-18, 1-19.

Salway, A.A. 1976. Statistical Estimation and Prediction of Avalanche Activity from

Meteorological Data for the Rogers Pass Area of BC. PhD Dissertation, University of

British Columbia, Vancouver, BC. 114 pp.

Schaerer, P.A., 1987. Avalanche Accidents in Canada III. A Selection of Case Histories

1978-1984, National Research Council of Canada, Institute for Research in

Construction, Paper No. 1468, 138 pp.

Schaerer, P.A. 1981. Avalanches. Handbook of Snow: Principles, Processes, Management

and Use, Edited by D.M. Gray and D.H. Male, Pergamon, 475-516.

Schaerer, P.A. 1989. Evaluation of the shovel shear test. Proceedings of the 1988

International Snow Science Workshop at Whistler, BC, 274-276.

Schaerer, P.A. 1991. Investigations of in-situ tests for shear strength of snow.

Unpublished, 29 pp.

Schleiss, V.G. and W.E. Schleiss. 1970. Avalanche hazard evaluation and forecast,

Rogers Pass, Glacier National Park. Ice Engineering and Avalanche Hazard

Forecasting and Control, National Research Council of Canada, Technical

Memorandum 98, 115-121.

241

Schweizer, J. 1991. Dry slab avalanches triggered by skiers. Proceedings of the

International Snow Science Workshop in Bigfork, Montana, 1990. ISSW '90

Committee, P.O. Box 372, Bigfork, Montana 59911, 307-309.

Schweizer, J. 1993. The influence of the layered character of snow cover on the triggering

of slab avalanches. Annals of Glaciology 18, 193-198.

Schweizer, J., M. Schneebeli, C. Fierz and P.M.B. Föhn. In press. Snow mechanics and

avalanche formation: Field experiments onto the dynamic response of the snow cover.

Surveys in Geophysics, 1-13.

Schweizer, J and P.M.B. Föhn. 1995. Two expert systems to forecast the avalanche

hazard for a given region. Proceedings of the 1994 International Snow Science

Workshop in Snowbird, Utah. ISSW '94, P.O. Box 49, Snowbird, Utah, USA,

295-309.

Schweizer, J., C. Camponovo, C. Fierz and P.M.B. Föhn. 1995. Skier-triggered slab

avalanche release - some practical implications. Proceedings of the International

Symposium at Chamonix, 30 May - 3 June 1995: The Contribution of Scientific

Research to Snow, Ice and Avalanche Safety. Association Nationale pour l'Etude de la

Neige et des Avalanches. Grenoble.

Schweizer, M. P.M.B. Föhn and J. Schweizer. 1994. Integrating neural networks and rule

based systems to build an avalanche forecasting system. Proceedings of the IASTED

International Conference: Artificial Intelligence, Expert Systems and Neuronal

Networks, Zurich, Switzerland.

Seligman, G.,1936. Snow Structure and Ski Fields. International Glaciological Society,

Cambridge, 555 pp.

Shapiro, S.S, M.B. Wilk and H.J. Chen. 1968. A comparative study of various tests for

normality. American Statistical Association Journal, 1343-1372.

Singh, H, 1980. A Finite Element Model for the Prediction of Dry Slab Avalanches, Ph.D.

Thesis, Colorado State University, Fort Collins, Colorado, 183 pp.

Smith, F.W. and J.O. Curtis, 1975. Stress analysis and failure prediction in avalanche

snowpacks, Snow Mechanics Symposium, Proceedings of the Grindewald Symposium,

242

International Association of Hydrological Sciences, Washington, D.C., Publication No.

114, 332-340.

Sommerfeld, R.A., 1973. Statistical problems in snow mechanics. U.S. Dept. of

Agriculture, Forest Service, General Technical Report RM-3, 29-36.

Sommerfeld, R.A., 1974. A weibull prediction of the tensile strength-volume relationship

of snow, Journal of Geophysical Research, 79(23), 3353-3356.

Sommerfeld, R.A. 1979. Accelerating strain preceding an avalanche. Journal of

Glaciology, 22(87), 402-404.

Sommerfeld, R.A., 1980. Statistical models of snow strength, Journal of Glaciology,

26(94), 217-223.

Sommerfeld, R.A. 1984. Instructions for using the 250 cm2 shear frame to evaluate the

strength of a buried snow surface. USDA Forest Service Research Note RM-446, 1-6.

Sommerfeld, R.A. and R.M. King, 1979. A Recommendation for the application of the

Roch Index for slab avalanche release, Journal of Glaciology, 22(87), 402-404.

Statsoft, 1994. Statistica for Windows: General Conventions and Statistics I. Statsoft Inc,

Tulsa, OK. 1718 pp.

Stethem, C. and R. Perla. 1980. Snow-slab studies at Whistler Mountain, British

Columbia, Canada. Journal of Glaciology 26(94), 85-91.

Stethem, C.J. and J.W. Tweedy, 1981. Field tests of snow stability. Proceedings of the

Avalanche Workshop in Vancouver, November 3-5, 1980, Edited by Canadian

Avalanche Committee, National Research Council of Canada, Technical Memorandum

No. 133, 52-60.

Stevens, J. E. Adams, X. Huo, J. Dent, J. Hicks and D. McCarty. 1995. Use of neural

networks in avalanche hazard forecasting. Proceedings of the 1994 International Snow

Science Workshop in Snowbird, Utah. ISSW '94, P.O. Box 49, Snowbird, Utah, USA,

327-340.

Topping, J. 1955. Errors of Observation and their Treatment. Chapman and Hall Ltd,

London. 119 pp.

243

Witmore, D, S.A. Burak, J. Malone and R.E. Davis. 1987. The Swiss rutschblock snow

stability evaluation test. Proceedings of International Snow Science Workshop at Lake

Tahoe, ISSW Workshop Committee, Homewood, California, 207-209.

Williams, K. and B. Armstrong. 1984. The Snowy Torrents: Avalanche Accidents in the

United States, 1972-79. Teton Bookshop, Box 1903, Jackson, Wyoming, 221 pp.

244

A ESTIMATING DENSITY FROM MICROSTRUCTUREAND RESISTANCE

A.1 Introduction

Estimated densities may be useful when densities have not been measured,

sometimes because the layer was too thin for the density sampler or the measurement was

omitted, possibly due to time constraints. This appendix outlines a method for estimating

density from observed microstructure and resistance.

A.2 Hand Hardness

The most widely used measure of resistance in Canada and internationally is “hand

hardness”. A fist, four finger tips, one finger tip, the blunt end of a pencil or a knife tip is

pushed horizontally into a snow layer while wearing gloves. The hand hardness is simply

the bluntest object that can be pushed into the snow with 10-15 N in Canada (CAA, 1989,

1995) or 50 N internationally (Colbeck and others, 1990). The levels of hard hardness are

abbreviated as F, 4F, 1F, P and K. Snow layers with resistance between the five major

levels of hand hardness can be qualified with a “+” or “-” sign, giving 15 levels: F-, F, F+,

4F-, 4F, 4F+, 1F-, 1F, 1F+, P-, P, P+, K-, K and K+ (CAA, 1995). Ice layers harder than

“knife” are labelled I for ice. Since the area of the objects being pushed into the snow does

not decrease proportionally from fist to knife, the hand hardness is an ordinal and not an

interval measure. However, to give a numeric scale suitable for graphing, the five major

levels are scaled geometrically based on doubling (NRCC/CAA, 1989; CAA, 1995). Using

1 for F, this results in a scale from 1 to 16 for the five major levels or 0.7 to 26.7 for the

15 major and minor levels as shown in Table A.1.

A.3 Mean Densities by Microstructure and Hand Hardness

Based on snow profiles done during the winters of 1993-95, density and hand

hardness were measured and microstructure observed for over 900 layers as summarized

in Table A.1. (Since substantial metamorphic and mechanical changes of snow often occur

within a day, the same level in the snowpack at the same location on different days is

245

246

Table A.1 Density of Layers Grouped by Hand Hardness and Microstructure

HandHardness

ScaledHand

HardnessRH

Precip.Particles1

Decomp./Fragmented

RoundedGrains

FacetedCrystals

DepthHoar

Crusts

N Mean±SD

N Mean±SD

N Mean±SD

N Mean±SD

N Mean±SD

N Mean±SD

F- 0.67 4 58±40

2 550

- - - - - - - -

F 1 46 86±31

70 112±29

9 179±45

11 167±49

5 212±44

- -

F+ 1.33 8 116±32

12 131±35

1 170 1 179 - - - -

4F- 1.7 - - 3 139±28

- - - - - - - -

4F 2 4 133±14

69 136±34

31 193±38

48 216±40

11

234±28

- -

4F+ 2.7 - - 14 156±32

7 196±30

4 222±14

2 305±7

1 148

1F- 3.3 - - 9 160±21

8 211±23

6 265±38

1 240 - -

1F 4 - - 40 164±34

95 207±37

83 262±36

14

260±59

- -

1F+ 5.3 - - 11 164±29

16 214±35

4 239±48

- - - -

P- 6.7 - - 6 212±29

35 248±37

3 272±16

- - - -

P 8 - - 5 217±58

122

271±42

50 294±43

3 290±36

3 222±28

P+ 10.7 - - 1 254 38 285±48

7 323±25

- - - -

K- 13.3 - - - - 1 302 - - - - - -

K 16 - - - - 4 308±57

4 317±52

1 270 16

259±54

K+ 26.7 - - - - - - - - - - 1 2761 excludes graupel and hail

considered to be a different layer.) Although the microstructure subclass (Colbeck and

others, 1990) was often recorded, the microstructure of layers is tabulated only by the

major class. Layers of graupel or hail are omitted from Table A.1 since there were only six

layers for which density and hand hardness were recorded and because the strength and

hence hardness of these layers are quite different from other subclasses of precipitation

particles (Section 6.2). The class of “wet grains” (which can include some types of dry

snow such as rounded polycrystals) is omitted since there were only two such layers for

which density and hand hardness were recorded. Although surface hoar is very important

to snow stability, it must be omitted because the layers were almost always thinner than

the diameter of the density sampler.

The mean densities based on three or more measurements for the remaining six

classes of microstructures are plotted against the scaled hand hardness in Figure A.1. For

these microstructures, there is a general increase in mean density with increasing hand

hardness. For most microstructures, the sample size is large enough that the density is

increased between minor levels of hand hardness. Precipitation particles show a rapid

increase in density with increasing hand hardness and a smooth transition to decomposed

and fragmented precipitation particles which is consistent with the common metamorphic

transition. The mean density of faceted crystals is close to that of depth hoar for particular

levels of hand hardness. The mean density of faceted crystals and depth hoar exceeds that

of rounded grains for most levels of hand hardness implying that, for a given density,

layers of rounded grains are generally harder than layers of faceted crystals and depth

hoar. Similarly for a given density, crusts are harder than layers of rounded grains, faceted

crystals and depth hoar which is indicative of the extensive bonding that is characteristic of

crusts. It is surprising that for densities between 175 and 220 kg/m3, layers of decomposed

and fragmented grains are harder than layers of rounded grains.

Linear equations of the form

ρ = A + B RH (A.1)

and logarithmic equations of the form

247

ρ = A + B ln RH (A.2)

where RH is the scaled hand hardness from Table A.1 and A and B are empirical constants

are fitted to the individual points for each major microstructure class. The empirical

constants along with the coefficient of determination, R2, and the standard error of

estimation, s, are given in Table A.2. Standard errors of estimation range between 32 and

49 kg/m3. Such high variability is not surprising since the regressions are based on a

ordinal measure of hardness.

Using R2 as a measure of fit, the linear equation (Eq. A.1) fits the data for layers of

rounded grains best and the logarithmic equation (Eq. A.2) best fits the data for the other

five microstructures. Since the linear equation only fits the data for rounded grains slightly

better than the logarithmic equation, Eq. A.2 is shown in Figure A.1 for the six

microstructures. Since Equation A.2, which is based on a logarithm of a geometric

sequence, best fits the data for five of the six microstructures, the merit of the geometric

scaling of hand hardness for estimating density is questionable. Nevertheless, the empirical

Figure A.1 Density by hand hardness for six common classes of microstructure.

248

equations from Table A.2 offer a means of estimating density from hand hardness for six

common classes of microstructure.

Table A.2 Regression Parameters for Estimating Density from Resistance andMicrostructure

Microstructure (Colbeck andothers, 1990)

No. ofLayers

Regression ρ = A + B RH Regression ρ = A +B ln(RH)

A B R2 p s A B R2 p s

PrecipitationParticles1 (1)

62 48 39 0.15 0.002 32 87 60 0.19 <10-3 32

Decomposed &Fragmented (2)

243 102 15 0.39 <10-6 33 109 43 0.40 <10-6 33

RoundedGrains (3)

367 167 12 0.41 <10-6 41 141 58 0.38 <10-6 42

FacetedCrystals (4)

220 204 11 0.37 <10-6 42 177 58 0.46 <10-6 39

Depth Hoar (5) 37 231 5 0.10 0.06 46 217 31 0.17 0.01 45

Crusts (9) 7 200 3 0.27 0.23 36 123 48 0.30 0.21 351 excludes graupel and hail

249

B ERROR ANALYSIS FOR STABILITY INDICES

B.1 Sources of Variability

Stability indices SN and SK depend on Daniels strength measured with the shear

frame, Σ∞, on slab density, ρ, on slab thickness in the start zone, h, and on slope

inclination, Ψ. In addition, SK depends on the estimate of penetration during skiing, PK.

In avalanche start zones, at least seven shear frame tests were usually made of the

failure plane, and coefficients of variation averaged 18% (Section 4.3), implying a

standard error of 7% of the mean strength.

Mean slab density was either measured once with a vertical density profile or one or

more times with a core sampler (Section 3.4). Coefficients of variation for mean slab

densities based on density profiles at a given site are typically 2-4% (Jamieson, 1989,

p. 67).

Mean slab thickness was usually measured to the nearest cm at two or more places

along the crown that appeared to be of average thickness. Since these measurements

rarely vary by more than 5% from the mean, the coefficient of variation is assumed to be

3%.

The slope inclination, Ψ, was usually measured at two or more locations that

appeared typical of the start zone. Since these typical values rarely vary by more than 2o,

the standard deviation of Ψ is approximately 1o and the coefficient of variation of cos Ψ or

sin Ψ is typically 1-2%.

The variability in PK depends on the regression on ρ30 in Section 6.8. For a given

value of ρ30, the standard deviation for PK is given by

s(PK|ρ30) = s(PK)(1-R2)½

(B.1)

From the data in Figure 6.21, R2 = 0.50 and s(PK) = 0.11 m, giving s(PK|ρ30) = 0.08 m

which is 28% of the mean value of PK.

250

B.2 Variability for Index SN

The stability index SN is given by

(B.2)SN =Σ∞ + σzzφ(σzz, Σ∞)

ρgh sin Ψ cos ΨHowever, φ ≅ 0 for the persistent layers that are central to this study, and the variability in

cos Ψ is only 1-2%, so the main sources of variability are Σ∞, ρ and h. Using the standard

formula for error propagation for uncorrelated measurements (Clifford, 1973), the

standard deviation for SN is

(B.3)s(SN) =

∂SN

∂Σ∞

2

s2(Σ∞) +

∂SN

∂ρ

2

s2(ρ) +

∂SN

∂h

2

s2(h)

1/2

Although Σ∞, ρ and h are likely correlated between sites since weak layers are generally

stronger under thicker, denser slabs, measurement variability probably obscures any

correlation for repeated measurements in a particular snow pit. Assuming that the

standard deviations of Σ∞, ρ and h are proportional to their means, Equation B.3 simplifies

to

(B.4)s(SN) = SN[v(Σ∞)2 + v(ρ)2 + v(h)2]1/2

where v(u) represents the coefficient of variation of a variable u and is the mean valueSN

of SN. Using the coefficients of variation for Σ∞, ρ and h from the previous section, the

coefficient of variation for SN is approximately 9% of its mean value.

The band of transitional stability 1 < SN < 1.5 used by Föhn (1987a) and Jamieson

and Johnston (1995a) can be interpreted as a one-sided confidence band above the critical

value of 1. Using SN = 1.5 and s(SN) = 0.09 for the standard error in the formula for the

confidence band

(B.5)SN − 1s(SN)

= t

gives t = 5.6. Since SN is usually based on at least seven shear frame tests, it has at least

six degrees of freedom, resulting in a 10-3 probability of a measured value of SN exceeding

1.5 when its true value is 1. Shear frame data with higher variability would result in lower

confidence attached to the safety margin 1 < SN < 1.5.

251

However, this approach to the safety margin is based on variability of measurements

such as shear strength within a snow pit and does not take into account the greater

variability within a start zone. The merit of SN and its safety margin really depends on the

proportion of prediction errors. However, such an approach requires more data than

presented in Section 6.5.

B.3 Variability for Index SK

Since the variability in slab density, ρ, and slab height, h, have a limited effect on the

variability of SN, the main sources of variability for stability index SK, are the measurement

of shear strength, Σ∞, and the estimate of skiing penetration, PK. Assuming Σ∞ and PK are

uncorrelated, the standard deviation for SK is

(B.6)s(SK) =

∂SK

∂Σ∞

2

s2(Σ∞) +

∂SK

∂PK

2

s2(PK)

1/2

where the partial derivatives of SK are

(B.7)∂SK

∂Σ∞= 1

σxz + ∆σ xz

and

(B.8)∂SK

∂PK= −2LΣ∞cos αmaxsin2αmaxsin(Ψ + αmax)

π(σxz + ∆σ xz)2(h − PK)2cos Ψusing the symbols introduced in Sections 2.4 and 6.8.

The variability of SK depends strongly on the term (h-PK)2 in the denominator of

Equation B.8. As the skiing penetration, PK, approaches the slab thickness, h, the term

(h-PK)2 approaches zero causing potentially unlimited variability for low values of SK.

However, such unlimited variability for low values of SK can, at worst, cause some false

unstable results that do not have serious consequences (Section 6.7). For h-PK > 0.8 m,

the stress due to the slab, σxz dominates the stress due to the skier, ∆σ'xz (Figure 6.19),

causing SK to approach SN, the variability of which is discussed in Section B.2. Since the

merit of SK depends on its ability to discriminate between stable and unstable slabs, the

variability of SK is most important near its critical value which is expected to fall between

1 and 1.5. Since the partial derivatives ∂SK/∂Σ∞ and ∂SK/∂PK are not simple functions of

252

Σ∞, ρ, h, PK and Ψ, the standard deviation of SK is estimated for the critical range.

However, there are only eight persistent skier-tested slabs for which 1 < SK < 1.5 in

Figure 6.22 whereas there are 36 persistent skier-tested slabs in the range 0.5 < SK < 2.

Using the larger set of 36 persistent skier-tested slabs, the mean values of the independent

variables are Σ∞= 1.27 kPa, ρ = 203 kg/m3, h = 0.64 m, PK = 0.24 m, Ψ = 38o, and

αmax = 46o (Föhn, 1987a) and the corresponding partial derivatives ∂SK/∂Σ∞ and ∂SK/∂PK

are approximately 1.03 kPa-1 and 1.20 m-1 resulting in an estimated standard deviation for

SK of 0.13 (Eq. B.5). Thus the regression estimate of PK causes the standard deviation of

SK to be approximately 50% greater than that of SN.

As was done with SN in Section B.2, the safety margin 1 < SK < 1.5 can be

interpreted as a one-sided confidence band above SK = 1. Replacing SN in Equation B.5 by

SK and using s(SK) = 0.13 for the standard error gives t = 3.85 implying a probability of

0.004 of a measured value of SK exceeding 1.5 when its true value is 1. However, this

approach to the safety margin is based on variability of measurements such as shear

strength within a snow pit and does not take into account the greater variability within a

start zone. The merit of SK and its safety margin really depends on the proportion of

prediction errors (Section 6.8).

253

C EXAMPLE OF FIELD NOTES

An example of two facing pages of field notes are shown in Figure C.1. Field notes

are made on specially prepared field books with pages of water resistant paper. In the

heading at the top of the pages, the weather, location, equipment and weak layer are

described in lines 1-4, respectively. Symbols for weather (CAA, 1995) show the sky was

overcast and snow was falling at less than 1 cm per hour when observations started at

1055.

The profile is recorded on the left page. Layer boundaries, in cm, are recorded in the

column headed H. In this example, boundaries are measured vertically upwards from the

ground. Hand hardness for the layers are recorded in the column headed R. These show

the hardest layers between 100 and 144 cm above the ground. The weak layer that failed

consisted of 2 mm faceted crystals (�) and 2-3 mm surface hoar (V). To save time

because of deteriorating weather, grain form (F) and size (D) were only recorded for the

substratum (called the bed surface), the weak layer that failed, and the surface hoar layer

buried on 10 February that did not fail.

The right page shows notes of a density profile (for calculating slab weight per unit

area), shear frame tests and rutschblock tests. In subsequent winters, densities were

measured layer-by-layer as described in Section 3.4. Eight shear frame tests were done

with a 0.025 m2 frame. Pull forces at failure ranged from 6.8 to 8.0 kg-force. Fracture

surfaces were all planar (marked "C" for "clean"). Only one rutschblock test was done

because of deteriorating weather. Ski penetration after gently stepping onto previously

undisturbed snow (SP) was 25 cm, and 36 cm after two jumps in the same place (JPx2).

After the skier moved onto the rutschblock column, the top 55 cm of the column slid on

loading step 6. The remaining 45 cm (down to 100 cm from the surface) did not displace

after repeated jumps, resulting in a score of 7.

Although redrafted, these are the field notes for the observations at the one-day-old

slab avalanche on Mt. Albreda described in Section 8.2.

254

255


Recommended