Вы находитесь на странице: 1из 487

University of Wollongong

Research Online
University of Wollongong Thesis Collection University of Wollongong Thesis Collections

1994

Engineering geologic factors influencing the


stability of slopes in the northern Illawarra region
Mohammad Hossein Ghobadi
University of Wollongong

Recommended Citation
Ghobadi, Mohammad Hossein, Engineering geologic factors influencing the stability of slopes in the northern Illawarra region, Doctor
of Philosophy thesis, Department of Civil and Mining Engineering, University of Wollongong, 1994. http://ro.uow.edu.au/theses/
1244

Research Online is the open access institutional repository for the


University of Wollongong. For further information contact the UOW
Library: research-pubs@uow.edu.au
ENGINEERING GEOLOGIC FACTORS INFLUENCING THE STABILITY OF SLOPES
IN THE NORTHERN ELLAWARRA REGION

A thesis submitted in fulfilment of the


requirements for the award of the degree of

DOCTOR OF PHILOSOPHY

from

THE UNIVERSITY OF WOLLONGONG

NEW SOUTH WALES, AUSTRALIA

by

M O H A M M A D HOSSEIN GHOBADI

(B.Sc in Geology; Ferdosi Uni. Mashad, Iran)


(M.Sc in Eng. Geology; Tarbiat modarres Uni. Tehran, Iran)

DEPARTMENTS OF GEOLOGY A N D CIVIL A N D MINING ENGINEERING


1994
Except where otherwise acknowledged, this thesis represents the author's original
research and the material included has not been submitted for a higher degree to any
others institution.

M o h a m m a d Hossein Ghobadi
ABSTRACT

This thesis is concerned with understanding the engineering geologic factors influencing

the stability of slopes along a coastal escarpment in the northern Illawarra region. The

particular area covered by this study lies between Coalcliff and Clifton where a number

of known or visible areas of slope instability are present. Slope instability in the study

area is a function of the stratigraphy, structural geology, petrology, geomorphology,

climate and mechanical properties of the rock and soil.

In the northern Illawarra region the essentially flat-lying lower Narrabeen Group

conformably overlies the upper Illawarra Coal Measures and the strata consist of

repeated beds of sandstone, shale, claystone and coal seams. The lower Narrabeen

Group consists of thick sequences of weak fine-grained rocks which are rather more

easily eroded than the associated sandstone strata and hence relatively rapid rates of

recession occur. Undermining along this the contacts between claystone and sandstone

reduces the support for the overlying vertically-jointed sandstone and eventually leads

to stabs falling off along the vertical joint faces. Thin marker beds (coal seams) in the

Illawarra Coal Measures and sandstone beds in the Narrabeen Group commonly act as

aquifers, with claystone beds acting as aquitards. -Slope instability is usually related to

the presence of the aquifers which are the source of high pore water pressures:- Perched

water tables have been found to be quite common in the study area.

-The topography along the escarpment is mainly steep and highly irregular because of

past and present marine and fluvial erosion:- Large colluvial deposits have accumulated

at the base of the steeper slopes. Generally, colluvial deposits are clay-rich containing

abundant mixed-layer clay and smectite (montmorillonite).

-Based on the petrological study, the Narrabeen group was derived from the New

England Fold Belt to the north and consists predominantly of volcanic detritus. The

volcanic detritus is present in both the sandstone and shale units either in form of

detrital grains of volcanic rock or as fine volcanic ash- During post-depositional


alteration and diagenesis the original volcanic glass in the ash and matrix of larger

grains has devitrified to produce smectite clays. These clays not only cause swelling

and shrinkage near the surface as a response to wetting and drying, but also reduce the

permeability of the near the surface rock mass. This latter factor increases the aqueous

pore pressures and hence increases the likelihood of surficial mass movement of both

the rock mass and the adjacent talus deposits.

Based on X-ray diffraction, carbonates are mostly rare in the talus deposits. The natural

reduction in carbonate cement due to weathering is one a cause for talus slope

instability in the Illawarra area.

The high horizontal stress environment known to exist in the Illawarra area is an

important factor which also influences slope failure. The resulting joint strike maxima

for the lower Narrabeen Group show that the most prominent joint set exposed at the

surface, with a direction between 005 and 025, has a significant effect on slope

stability in the study area.

Fracture permeability is also the most important feature of groundwater movements with

it most of the fractures occurring in areas of stress relief near the face of the

escarpment. It is quite obvious from studying the rainfall figures and periods of

prevalence of landslides that the most unstable periods are those when the rainfall is

above 400 mm per month.

A significant decrease in durability was found to accompany changes in mineralogy and

an increase in weathering from fresh to weathered rocks. Moderately and highly

weathered claystone and shale in the Narrabeen Group rocks have low to very low

durability; it is dependent on their mineralogy, and especially on the type and quantity

of clay minerals present. Claystone samples interbedded in the Bulgo Sandstone also

show very low durability. In contrast, claystone interbedded in the Scarborough

Sandstone shows a medium durability whereas claystone in the Coal Cliff Sandstone has

a high durability. The differences in the behaviour of samples is that slake durability

is sensitive to the abundance of clay minerals as opposed to carbonate in these samples.


Claystone interbeds in the Bulgo Sandstone and the highly weathered Stanwell Park

Claystone both have very low durability. This has a significant effect on slope stability

in the Bulgo Sandstone especially where the Stanwell Park Claystone acts as the

bedrock for the talus mantle between Clifton and Stanwell Park.

The Wombarra Shale and Stanwell Park Claystone, two units of the Narrabeen Group,

appear to dominate the study area as being the units most prone to instability problems.

Failure surfaces of landslides are located at or near the base of highly weathered shale

or claystone sequences.

A significant decrease in strength was also found to occur with an increase in

weathering from fresh to weathered rocks. The geotechnical properties of the talus

most related to its stability, are clay content, plasticity index and residual friction angl

These parameters and the angle of natural slopes show the talus is unstable in the long-

term at slopes above 10-12.

Man's construction activities have also caused some landslides in the northern Illawarra,

especially along the excavations for the railway and road. Two main transport routes,

the Illawarra Railway and Lawrence Hargrave Drive, pass through the study area.

Along Lawrence Hargrave Drive major movement in gently sloping land has been

triggered by high pore-water pressures in highly weathered Wombarra Shale covered by

a talus mantle. Increased urban development has and will continue to complicate the

issue in the future. Seven landslides have been detailed in this thesis. The majority

of these have or are presently undergoing block type movements at creep rate. Detailed

geotechnical investigations with survey monitoring is often necessary to identify these

failures.

The area has also been extensively mined for coal, resulting in minor subsidence. This

has usually caused fracturing of the rock strata and opening of the joint system which

have increased water ingress, resulting in higher subsurface flows and altered

groundwater regimes. Based on observations, mine subsidence has been one

contributory factor to slope instability in the northern Illawarra.


ACKNOWLEDGMENTS

The work represented in this thesis was carried out under the supervision of Associate

Professors B.G. Jones and R.N. Chowdhury. I am indebted for their constant

encouragement, guidance and discussions.

Facilities for carrying out the investigations were provided by the Departments of

Geology and Civil and Mining Engineering at the University of Wollongong.

My sincere thanks are due to Dr D. Titheridge and Mr P. Lamb (Kembla Coal & Coke

Pty Ltd) for their help in preparation of rock samples in the field and access to

photographs and comparative data. Special thanks are due to Messrs J. Peterson from

the Department of Main Roads (Wollongong office) and H.D. Christie from State Rail

Authority of New South Wales (Geotechnical Section) for their assistance.

I am thankful to Dr J.V. Hamel (Hamel Geotechnical Consultants, USA) for offering

valuable suggestions and advice. I am also grateful to the staff of the Departments of

Geology and Civil and Mining Engineering, and my friends for their support and co-

operation.

Financial assistance was received for this study from a Postgraduate Scholarship

provided by the Islamic Republic of Iran. I wish to acknowledge this financial

assistance in preparing this thesis.

Consistent encouragement and deep interest shown by family members, especially my

wife, have provided necessary inspiration. I appreciate their thoughtfulness in bearing

with me and the inconvenience due to my constant occupation with this work in the

final stages.
CONTENTS

CHAPTER 1

1.1 INTRODUCTION 1

1.2 AIMS 3

1.3 ILLAWARRA REGION 5

1.4 PREVIOUS WORK 6

1.5 STUDY METHODS 8

1.6 MASS MOVEMENT 10

1.6.1 LANDSLIDE TERMINOLOGY 10

1.6.2 TYPE OF MASS MOVEMENT 10

1.6.3 FACTORS CAUSING MASS MOVEMENTS 14

1.7 MAIN CAUSES OF LANDSLIDES IN THE 18


ILLAWARRA AREA

CHAPTER 2
GEOLOGICAL AND GEOGRAPHICAL REVIEW
OF THE ILLAWARRA REGION
2.1 REGIONAL GEOLOGY 23
2.2 STRUCTURAL GEOLO< 24

2.3 ILLAWARRA AREA 26

2.3.1 GEOMORPHOLOGY 26
2.3.2 CLIMATE 28

2.3.3 GEOLOGY 29
2.3.4 STRATIGRAPHY 30
2.3.4.1 Shoalhaven Group 30
2.3.4.2 Illawarra Coal Measures 30
2.3.4.3 Narrabeen Group 31
2.3.4.4 Coal Cliff Sandstone 32
2.3.4.5 W o m b a r r a Shale 32
2.3.4.6 Scarborough Sandstone 33

2.3.4.7 Stanwell Park Claystone 33

2.3.4.8 Bulgo Sandstone 34

2.3.4.9 Bald Hill Claystone 34

2.3.4.10 Garie Formation 35

2.3.4.11 Newport Formation 35

2.3.5 POST-NARRABEEN UNITS 35

2.3.5.1 Hawkesbury Sandstone 35

2.3.5.2 Wianamatta Group 36

2.3.5.3 Igneous Rocks 36

2.3.5.4 Tertiary and Quaternary Deposits 37

2.3.6 STRUCTURAL G E O L O G Y 37

2.3.6.1 Folds 37

2.3.6.2 Faults 38

2.3.6.3 Joints 39

2.7 S T R E S S FIELDS 40

CHAPTER 3
GEOLOGY OF THE UPPER COAL MEASURES AND LOWER NARRABEEN GROUP
IN THE SCARBOROUGH-STANWELL PARK AREA
3.1 INTRODUCTION 41
3.2 ILLAWARRA COAL MEASURES 42
3.2.1 UPPER ILLAWARRA COAL MEASURES 43
(SYDNEY SUB-GROUP)
3.2.2 WILTON FORMATION 43
3.2.3 TONGARRA COAL 44
3.2.4 BARGO CLAYSTONE 44
3.2.5 DARKES FOREST SANDSTONE 44
3.2.6 ALLANS CREEK FORMATION 45
3.2.7 KEMBLA SANDSTONE A*
3.2.8 WONGAWILLI COAL 46
3.2.9 ECKERSLEY FORMATION 46
3.2.10 BULLI COAL 47
3.3 NARRABEEN GROUP 47
3.3.1 LOWER NARRABEEN GROUP 48
3.3.1.1 Coalcliff Sandstone (CSs) 48
3.3.1.2 Wombarra Shale (WSh) 49
3.3.1.3 Scarborough Sandstone (SSs) 50
3.3.1.4 Stanwell Park Claystone (SPC) 50
3.3.1.5 Bulgo Sandstone (BSs) 51
3.4 IGNEOUS ROCKS 52
3.5 TALUS 53
3.6 SEDIMENTARY STRUCTURES 54
3.6.1 SEDIMENTARY ENVIRONMENTS 54
3.7 SUBSURFACE GEOLOGICAL SEQUENCES AND 55
STRUCTURES RECOGNISED IN DRILL HOLES
3.8 GEOLOGY AND SEDIMENTARY STRUCTURE 55
3.8.1 UPPER ILLAWARRA COAL MEASURES 55
3.8.2 COAL CLIFF SANDSTONE 56
3.8.3 WOMBARRA SHALE 56
3.8.4 SCARBOROUGH SANDSTONE 57
3.8.5 STANWELL PARK CLAYSTONE 58
3.8.6 BULGO SANDSTONE 58
3.9 DISCUSSION 59

CHAPTER 4
PETROLOGY OF NARRABEEN GROUP SANDSTONE
4.1 INTRODUCTION 63
4.2 STUDY METHODS 63
3 MINERAL COMPOSITION 64
4.3.1 QUARTZ 64

4.3.2 FELDSPAR 65

4.3.3 ROCK FRAGMENTS 66

4.3.4 CHERT 66

4.3.5 MICA 66

4.3.6 ACCESSORY MINERALS 67

4.3.7 IRON OXIDES 67

4.3.8 CARBONATE 67

4.3.9 KAOLINITE 68

4.3.10 CLAY MATRIX 68

4.3.11 CEMENT 69

4.3.12 POROSITY 69

4.3.13 TEXTURE OF SANDSTONES 70

4.3.14 CLASSIFICATION OF SANDSTONE 70

4.4 PETROLOGICAL AND MINERALOGICAL ASPECTS 71


OF WEATHERING
4.5 MINERAL IDENTIFICATION USING X-RAY 74
DIFFRACTION
4.5.1 INTRODUCTION 74

4.5.2 AIM OF STUDY 74

4.5.3 METHOD OF STUDY 75

4.5.3.1 Sample Collection 75

4.5.3.3 Whole Rock Analysis 75

4.5.3.4 Clay Mineral Analysis 77

4.5.3.5 Normal or Untreated samples 7g


4.5.3.6 Glycolation 7g

4.5.3.7 Heated Samples 79

4.6 RESULT OF X-RAY ANALYSIS 79


4.6.1 KAOLINITE 80
4.6.2 ILLITE 80
4.6.3 SMECTITE (MONTMORILLONITE) 81
4.6.4 MIXED - LAYER CLAYS 81
4.7 MUD ROCKS AND SANDSTONES 81
4.8 TALUS 83
4.8.1 COMPOSITION 83
4.9 INTERPRETATION OF THESE RESULTS 84
4.10 CLAY MINERAL STRUCTURE AND SLOPE 84
STABILITY
4.11 DISCUSSION 85
4.12 RELATIONSHIP BETWEEN PETROLOGY, SOURCE 88
AND SLOPE STABILITY

CHAPTER 5
STRUCTURAL GEOLOGY IN THE SLIP AREA
5.1 INTRODUCTION 91
5.2 STRUCTURAL FACTORS WHICH ARE IMPORTANT 91
IN SLOPE STABILITY
5.3 FAULTS IN THE SLIP AREA 94
5.3.1 HARBOUR FAULT 95
5.3.2 JETTY FAULT 95
5.3.3 CLIFTON FAULT 96
5.3.4 SCARBOROUGH FAULT 96
5.4 JOINTS IN THE SLIP AREA 96
5.4.1 JOINTS IN COAL 97
5.4.2 JOINTS IN THE NARRABEEN GROUP AND 98
HAWKESBURY SANDSTONE
5.5 THE IMPORTANCE OF FAULTS AND OTHER 101
THROUGH-GOING GEOLOGIC STRUCTURES
5.6 JOINTING AND TECTONIC FRACTURING OF ROCK 102
5.6.1 BEDDING 103
5.7 THE RELATIONSHD? BETWEEN JOINTS AND THE 104
ORIENTATION OF CLIFF FACE
5.7.1 STRESS RELIEF 104
5.8 THE RELATIONSHIP BETWEEN JOINTS AND 106
RATES OF EROSION OF STRATIGRAPHIC UNITS
IN DIFFERENT TYPES OF EXPOSURES
5.8.1 DIFFERENTIAL EROSION 107
5.9 SUMMARY AND CONCLUSION 110

CHAPTER 6
REVIEW OF THE ROLE OF GROUNDWATER, RAINFALL,
HYDROGEOLOGY AND EARTHQUAKES
6.1 GROUNDWATER 115
6.2 INTRINSIC PROPERTIES 115
6.2.1 POROSITY 116
6.2.2 PERMEABILITY 116
6.2.3 RELATIONSHIP BETWEEN POROSITY AND 117
PERMEABILITY
6.2.4 FRACTURE (SECONDARY) PERMEABILITY 117
6.3 POREWATER PRESSURE 118
6.4 CHANGES IN WATER CONTENT 119
6.5 EFFECTS OF SOLUTION 120
6.6 GROUNDWATER FLOW IN SLOPE STABILITY 120
PROBLEMS
6.7 SLOPES COVERED WITH LANDSLIDE DEBRIS 121
6.8 HIGH WATER PRESSURES IN THE ESCARPMENT 122
6.9 SPECIAL EFFECTS OF FAULTS ON THE 123
HYDROGEOLOGY OF SLOPES
6.10 HYDROGEOLOGICAL ASPECTS OF THE 124
ESCARPMENT IN THE STUDY AREA
6.11 RAINFALL AND ITS RELATIONSHIP TO 127
HYDROGEOLOGY
6.12 RAINFALL AND ITS RELATIONSHIP TO LAND 128
MOVEMENTS
6.12.1 THE CONCEPT OF THRESHOLDS 130
6.13 GROUNDWATER AND ITS RELATIONSHIP TO 131
LAND MOVEMENTS
6.14 SUMMARY AND CONCLUSION 132
6.15 EARTHQUAKES 135
6.15.1 EARTHQUAKES IN THE STUDY AREA 136
6.15.2 SECONDARY EFFECTS OF EARTHQUAKES 137
6.15.3 INTERPRETATION AND EFFECTS OF 137
EARTHQUAKES AND STRESS ENVIRONMENT

CHAPTER 7
ENGINEERING GEOLOGY
7.1 ENGINEERING PROPERTIES OF ROCKS IN THE 139
LOWER NARRABEEN GROUP
7.2 WEATHERING 140
7.2.1 ENVIRONMENTAL FACTORS CONTROLLING 140
ROCK WEATHERING
7.2.2 MINERAL HYDRATION 141
7.2.3 MINERAL SOLUTION 142
7.2.4 PROCESSES AND MECHANISMS OF WEATHERING 143
IN THE STUDY AREA
7.2.5 WEATHERING, STRENGTH AND LANDSLIDES 148
7.3 SLAKE DURABILITY TEST 149
7.3.1 INTRODUCTION 149
7.3.2 SLAKE DURABILITY 151
7.3.3 AIM OF STUDY 152
7.3.4 METHOD OF STUDY 152
7.3.4.1 Sample Collection 152
7.3.4.2 Sample Preparation 153
7.3.4.3 Procedure 153
7.3.4.4 Calculations 154
7.3.5 RESULTS 154
7.3.6 SLAKE DURABILITY CLASSIFICATION 156
7.3.7 STATIC (LONG-TERM) DURABILITY TESTING 156
7.3.8 CONCLUSIONS 157
7.4 POINT LOAD STRENGTH TEST 160
7.4.1 INTRODUCTION 160
7.4.2 THE AIM OF STUDY 161
7.4.3 METHOD OF STUDY 161
7.4.3.1 Sample Collection and Preparation 161
7.4.5 DIAMETRAL TESTS 162
7.4.6 AXIAL TESTS 162
7.4.7 IRREGULAR LUMP TESTS 162
7.4.8 CALCULATIONS 163
7.4.9 RESULTS 164
7.4.10 RELATIONSHIP BETWEEN POINT LOAD 165
STRENGTH INDEX AND UNIAXIAL COMPRESSIVE
STRENGTH
7.4.11 CONCLUSIONS 166
7.4.12 RELATIONSHIP BETWEEN UNIAXIAL 167
COMPRESSIVE STRENGTH (UCS) AND SLAKE
DURABILITY INDEX (SDI)
7.5 ROCK COMPOSITION IN RELATION TO 168
MECHANICAL PROPERTIES
7.5.1 ROCK COMPOSITION AND STRENGTH 168
7.5.2 ROCK COMPOSITION AND SLAKE DURABILITY 169
7.5.3 RELATIONSHIPS BETWEEN SLAKE DURABILITY, 172
ROCK STRENGTH AND WEATHERING IN
RELATION TO ROCK SLOPE AND TALUS FAILURE
ALONG THE NORTHERN ILLAWARRA COASTLINE
CHAPTER 8
SLOPE STABILITY IN THE NORTHERN ILLAWARRA
8.1 INTRODUCTION 177

8.2 SLOPE DEVELOPMENT PROCESS 178

8.3 RELATIVE IMPORTANCE OF OTHER 180


GEOLOGICAL FACTORS FOR SLOPE FAILURES
A L O N G THE ILLAWARRA COASTLINE
8.4 ENGINEERING GEOLOGIC FAILURE MODELS FOR 182
SLOPE INSTABILITY ALONG THE NORTHERN
ILLAWARRA COASTLINE
8.5 TYPES OF SLOPE INSTABILITY 184

8.5.1 ROCKFALLS AND TOPPLING 184

8.5.2 SHALLOW DEBRIS SLIDES 185

8.5.3 DEBRIS FLOWS 185

8.5.4 DEEP-SEATED SLUMP-EARTH FLOWS 185

8.5.5 CREEP 186

8.6 FAILURE OF TALUS SLOPES 186

8.6.1 INTRODUCTION 186

8.6.2 ORIGIN OF THE TALUS 187

8.6.3 PARENT MATERIAL OF TALUS 187

8.6.4 CLAY MINERAL ANALYSES 188

8.6.5 GEOTECHNICAL PROPERTIES OF TALUS 189

8.6.6 TEST RESULTS 190

8.6.6.1 Strength Characteristics of the Talus Matrix 190

8.6.6.2 Correlation of Engineering Indices and Properties 191

8.6.7 CONCLUSION 193

8.7 CASE STUDIES 195

8.7.1 SITE 1 CLIFTON EARTH SLUMP 196

8.7.1.1 Location 196

8.7.1.2 Geology 196

8.7.1.3 Description of the slump 197


8.7.1.4 Geotechnical properties of the talus 198
198
8.7.1.5 Conclusions

8.7.2 SITE 2 MORONGA PARK SLUMP-EARTH FLOW 199


199
8.7.2.1 Location
19
8.7.2.2 Geology 9

8.7.2.3 Description of the slump-earth flow 200

8.7.2.4 Geotechnical properties of the talus 201

8.7.2.5 Conclusions 202

8.7.3 SITE 3 SOUTHERN AMPHITHEATRE COMPLEX 203


LANDSLIDE (GRABEN A)

8.7.3.1 Introduction 203

8.7.3.2 Location 203

8.7.3.3 Geology 203

8.7.3.4 Description of the slide 205

8.7.4 SITE 4 NORTHERN AMPHITHEATRE COMPLEX 206


LANDSLIDE (GRABEN B)

8.7.4.1 Location 206

8.7.4.2 Geology 206

8.7.4.3 Description of the slide 206

8.7.4.4 Conclusion 208

8.7.5 SITE 5 JETTY ROCK SLUMP 210

8.7.5.1 Location 210

8.7.5.2 Geology 210

8.7.5.3 Description of rock slump 211

8.7.5.4 Geotechnical properties of the talus 211

8.7.5.5 Conclusion 212

8.7.6 SITE 6 HARBOUR SLUMP 212

8.7.6.1 Location 212

8.7.6.2 Geology 212


8.7.6.3 Description of the slump 213

8.7.6.4 Geotechnical properties of the talus 213

8.7.6.5 Conclusion 214

8.7.7 SITE 7 COALCLIFF SLUMP 214

8.7.7.1 Location 214

8.7.7.2 Geology 214

8.7.7.3 Description of the slump 215

8.7.7.4 Geotechnical properties of the talus and Stanwell Park 2


Claystone
8.7.7.5 Conclusion 217

8.8 SURFACE SURVEY RESULTS 217

8.9 FAILURE OF ROCK 218

8.9.1 INTRODUCTION 218

8.9.2 MECHANICS OF ROCK FAILURES 219

8.9.3 EFFECTS OF WEATHERING AND JOINTING 220

8.9.4 CREEP 222


8.9.5 ROCKFALL AND TOPPLING ALONG THE 223
L A W R E N C E HARGRAVE DRIVE
8.9.6 ROCKFALLS AND TOPPLING ALONG THE 224
COASTLINE

8.9.7 CONCLUSIONS 224

8.10 TREATMENT, STABILISATION AND PREVENTION 226

CHAPTER 9
S U M M A R Y AND CONCLUSIONS
9.1 INTRODUCTION 231

9.2 STRATIGRAPHY 232

9.3 PETROLOGY 232

9.4 STRUCTURE 233

WATER 234
9.6 GEOTECHNICAL PROPERTIES OF ROCK AND 234
TALUS
9.7 SLOPE DEVELOPMENT PROCESS 235
9.8 REMEDIAL WORKS 237
9.9 CONCLUSIONS 238

REFERENCES 241
FIGURES
FIGURES TO CHAPTER 1 FIGURE 1.1 - FIGURE 1.12
FIGURES TO CHAPTER 2 FIGURE 2.1 - FIGURE 2.12
FIGURES TO CHAPTER 3 FIGURE 3.1 - FIGURE 3.23
FIGURES TO CHAPTER 4 FIGURE 4.1 - FIGURE 4.16
FIGURES TO CHAPTER 5 FIGURE 5.1 - FIGURE 5.25
FIGURES TO CHAPTER 6 FIGURE 6.1 - FIGURE 6.15
FIGURES TO CHAPTER 7 FIGURE 7.1 - FIGURE 7.33
FIGURES TO CHAPTER 8 FIGURE 8.1 - FIGURE 8.55

TABLES
TABLES TO CHAPTERS 1-3 TABLE 1.1 - TABLE 3.1
TABLES TO CHAPTER 4 TABLE 4.1 - TABLE 4.25
TABLES TO CHAPTERS 5-6 TABLE 5.1 - TABLE 6.1

TABLES TO CHAPTER 7 TABLE 7.1 - TABLE 7.29


TABLES TO CHAPTER 8 TABLE 8.1 - TABLE 8.3

APPENDICES
APPENDIX TO CHAPTER 7 TABLE 1 - TABLE 16
APPENDIX TO CHAPTER 8 EXAMPLES OF SHEAR TEST
APPENDIX - SAMPLE LOCATIONS
GLOSSARY
CHAPTER 1
INTRODUCTION
1.1 INTRODUCTION
It is commonly accepted that geological understanding can provide the basis for

predicting slope stability in any area, and considerable research effort has been directed

towards understanding and quantifying the influence of geological factors such as

structure, stratigraphy, sedimentology, petrography, weathering, groundwater,

geomorphology and earthquakes on slope stability (Guadogno et aL, 1994; Cascini et

al, 1994).

Many types of slope movement and landslides occur in residual soils, talus and

colluvium. A proper understanding of these landslides also requires an understanding

of geological processes, local geological details and the role of geological discontinuities.

Of all the geologic factors influencing the stability of rock slopes, there is little doubt

that joints, bedding planes, faults and shear zones, and intersections of such structures,

are the most significant. Chemical alteration of the surrounding rock and presence of

clay gouge are also commonly associated with fault and shear zones. The presence of

clay gouge adjacent to the polished or smooth rock surfaces of faults is often associated

with the unusually low shear strength for soil-rock surfaces.

Stratigraphy is the basis for most rock slope design procedures (Philbrick, 1960).

Therefore, one of the first steps in the investigation of a slope should be the

determination of the stratigraphy of the area. Particular attention should be directed

towards recognition and description of thick sequences of weak rocks (claystone,

siltstone and shale) that are very important in slope development, and have a notoriously

high landslide potential (Kelley, 1971; Winters, 1972). They may have also contributed

to deep-seated ancient landsliding. Thin marker beds, such as fossiliferous and non-

fossiliferous limestone, coal and clay seams, and carbonaceous shale are extremely
i
important in determining the extent of previous landsliding. Coal seams commonly act
2

as aquifers and their underclays (claystone) often contribute to failure surface

development in colluvial slopes. For example, coal beds in both the Newcastle and

Illawarra Coal Measures commonly act as aquifers, with claystone beds acting as

aquitards. Landsliding is usually related to the presence of the aquifers which are a

source of high pore water pressures. The shear strength along any potential surface of

sliding decreases as the pore water pressure increases.

The distribution and properties of rocks in an area is largely controlled by conditions

at the time of their original deposition. Depositional environments in the Illawarra

region range from fluvial to deltaic, intertidal facies, marginal marine and marine. Each

environment has a characteristic association of sedimentary facies with a specific range

of properties, and this stratigraphy affects slope stability in each area.

Another geological factor affecting slope stability is petrography; percentages of the

principal rock-forming minerals and cements provide a key index to mechanical

performance. Quartz percentage is an important characteristic of sedimentary rocks,

particularly of siltstone and shale, and is an indicator of their strength and abrasiveness

The total clay content, and at least the approximate percentage of clay mineral types,

are useful indicators of the potentially plastic and swelling behaviour of shale (Franklin

and Dusseault, 1989). The engineering behaviour and especially the residual strength

of colluvium derived from argillaceous rocks also depends on its (original and

weathered) petrography (Hamel, 1980).

Weathering, both mechanical and chemical, gradually disturbs the cohesion of rocks.

In many landslide events, chemical alterations, such as hydration and ion exchange in

clays, are thought to have contributed to the triggering of landslides (Zaruba and

Mencel, 1969).

Geomorphology is concerned with the nature and origin of landforms and the study of

processes of landform development. Geomorphological information is useful in


3

understanding the complex phenomena and many interacting factors which control

landform evolution and shape, and its relationship to slope behaviour.

Regional and local groundwater conditions are often very critical to slope stability.

Slope failures are often associated with high groundwater levels following prolonged or

intense rainfall. The influence of groundwater on stability can operate in two ways:

(1) excessive seepage; and

(2) excessive pore water pressure.

If the rock mass is very tight, or if the soil mass is impermeable, water pressure can

exist when groundwater seepage is low. Alternatively, if abundant water is available

and the rock mass is open or the soil mass is permeable, then substantial seepage of

water can occur under quite low hydrostatic pressure.

The manner in which water pressure enters into the consideration of stability is made

clear by an understanding of the concept of effective stress. Water pressure reduces the

normal effective stress resulting in shear strength decrease; consequently draining a rock

or soil mass will reduce the water pressure and increase the shear strength.

Almost every earthquake in mountainous country, strong enough to be reported in the

literature, has caused at least some landslide and/or rock fall. Unless these slides or

falls damage structures or seriously block transportation, they often go unreported. For

example, in northeastern Honshu (Japan), an earthquake in 1978 (magnitude 7.4) caused

thousands of slides, most of them less than 10 cubic metres in volume, but some big

rockfall occurred as well.

1.2 AIMS

Comprehensive evaluation of the geological aspects of slope stability in the area

between Coalcliff and Clifton (Fig. 1.4) is the primary and main objective of the

work reported in this thesis. Therefore, in accordance with the approved

programme, about 90 percent of the total research effort has been devoted to the
4

first two of the following four aims. T h e remaining 10 percent of total research

effort has been devoted to the third and fourth aims in the list.

The four aims, with relative weighting shown in parentheses at the end of each aim, are

as follows.

Maior A i m s

(1) T o conduct a literature review on the mechanisms of landslides and factors

affecting slope stability. In particular, the influence of geological factors, such as

stratigraphy, petrography, structure, weathering and groundwater, which cause talus

failure and rock failure (toppling and rockfall), is to be investigated. (20 percent

of work)

(2) T o test and develop an understanding of the effects of weathering and slaking on

mechanical and mineralogical properties of fresh and weathered rocks in the

Narrabeen Group between Coalcliff and Clifton, and to discuss the implications

of these processes on slope stability. (70 percent of work)

Minor A i m s

(3) T o determine the role of geotechnical properties of talus which control the stability

of slopes at shallow depths within talus. The third aim includes an integration of

currently available and n e w data from talus deposits in the Northern Illawarra with

the previously defined geological constrains provided in the second aim. It should

be noted that facilities and funds were not available to carry out appropriate

drilling necessary for any n e w stability analyses. Therefore, work concerning

properties of talus was a relatively minor aim of the thesis. (8 percent of work)

(4) T o briefly review remedial measures that could alleviate the problem of land

instability in the Illawarra region, and to identify the relevance of the obtained

geological data in defining future directions for research. (2 percent of work)

Geotechnical stability analyses were outside the scope of the thesis due to the lack of

funding to carry out a significant amount of expensive drilling in very steep and
5

inaccessible terrain in the northern Illawarra region. Moreover, preparation of

engineering geological maps or stability maps was also outside the scope of this thesis.

Such maps are being developed for associated doctoral programme at the University of

Wollongong using a Geographical Information System (GIS) package (Flentje, 1996).

1.3 ILLAWARRA REGION


In New South Wales, four main regions of landsliding have been identified: Illawarra,

Campbelltown-Picton, Gosford-Wyong (south, southwest and north of the Sydney,

respectively) and Lismore (Fig. 1.1). The Illawarra is located in the Greater

Wollongong area, as shown in (Figs 1.2 and 1.3). The specific area chosen for study

is between Coledale and Coalcliff as illustrated in (Fig. 1.4). The Greater Wollongong

area has long been recognised as a region of major landslide activity (e.g Hanlon, 1952,

1953, 1958; Walker, 1960; Bowman, 1972; Chowdhury, 1976; Young, 1977, 1978).

This activity is directly related to the geology, geological history and geomorphology

of the area.

The 300 m high escarpment consists of flat-lying Permo-Triassic volcaniclastic coal

measures plus fluviatile sequences capped by a well cemented quartz sandstone (Fig.

1.5). Flat-lying interbedded strong and weak sedimentary rocks in the Illawarra region

have been acted upon by erosion, stress relief, weathering, creeping and sliding

processes to produce masses of marginally stable colluvial soil and zones of potentially

unstable rock masses on many of the steep hillsides that are common to the area.

The traditional modes of failure, namely plane failure, wedge failure and rotational

failure often occur, and the type of failure is related to the rock structure and

weathering which play an important role (Fig. 1.6). Toppling failures, rockfalls and

failures of talus or colluvium have occurred in the past and continue to occur now on

moderate to steep slopes in the Illawarra area. Many of the rock slopes are relatively
6

simple in their form and shape. In the vast majority of places the bedding is nearly

horizontal and jointing in the rock is approximately vertical.

Slope stability evaluations in the Illawarra area are interdisciplinary geotechnical

endeavours requiring concepts from engineering geology, soil mechanics and rock

mechanics. Of these disciplines, engineering geology is probably the most important.

Reliable evaluations of slope stability must begin with an understanding of regional and

site geology and the geologic processes which formed the site and continue to act upon

it. Once this level of geologic understanding is reached, slope behaviour can often be

assessed on the basis of judgement aided by simple analyses, experience and precedence.

Where detailed stability analyses are required, the above-mentioned geologic

understanding is mandatory for development of appropriate geotechnical models.

Many instability problems in this area relate to failures of talus or colluvium, which is

derived from older failures on the escarpment. Studies concerned with this type of

slope stability, also need to deal with failure of the escarpment and characteristics of

materials which affect these failures. These materials range from fresh to completely

weathered claystone and sandstone.

This thesis has focused on the talus and rock slopes, and an understanding of the

engineering geological features of the area. However, considerably more attention must

be paid to the environmental geology of Illawarra (including weathering, excavation,

drainage, storms and rainfall intensity) because these aspect need comprehensive

understanding for improved assessment of the problems associated with various types

of mass movement.

1.4 PREVIOUS WORK

Experience in the Illawarra area has shown that slope instability is an old problem that

can be worsened by disturbance and development (Fig. 1.1). Moreover, slope instability

has had disastrous effects on urban area as well as road and rail routes. As long ago
7

as 1890, W . Shellshear published a paper entitled "on treatment of slip on the Illawarra

railway at Stanwell Park ". This was at the time when the railway line occupied the

position of the present main road (Lawrence Hargrave Drive).

Hanlon (1952, 1953, 1958) discussed in some detail the causes of rock failures in the

Wollongong area and suggested various remedial measures. Hanlon's studies focussed

on the role of lithology and structure in determining the stability of escarpment slopes.

Walker (1960) mapped soils "developed on debris avalanche deposits" in a continuous

zone right along the escarpment from Wollongong to Nowra. Since then, m a n y authors

have continued to research the problems. B o w m a n (1972) made a thorough descriptive

investigation of the natural slope stability in the Greater Wollongong area and although

he was mainly concerned with slope stability in relation to the residential subdivisions,

he highlighted the importance of water and jointing in relation to the stability of the

escarpment. Young (1976) studied talus-mantled deposits, which she referred to as

'taluvium'. She also mentioned the escarpment evolution and assessed the influence of

local climatic variation on slope stability near Wollongong (1977, 1978). Chowdhury

(1976) highlighted the importance of the mechanism of progressive slope failures based

on the decrease in shear strength due to weathering and stress release. Evans (1978)

studied "time dependent factors influencing the rock slope stability of the Illawarra

escarpment". His work was concerned with weathering and creep, but he also described

a differential settlement mechanism which involves stress relaxation in the basal softer

rock (claystone) in relation to secondary toppling rock failures on the Illawarra

escarpment (Evans 1981). Chestnut (1981) mapped the Wollongong-Port Hacking area

with an emphasis on engineering geology and environmental geological hazards. Walker

et al. (1987) presented some of the results of a geotechnical study of the Coledale area

which is located on the Illawarra escarpment about 17 k m north of Wollongong.

M o r e recently Hutton et al. (1990) studied the landslide activity after the heavy rains

of April 1988 in the Coalcliff area of the northern Illawarra. They briefly described
8

landslides that have been prominent during the last few years along Lawrence Hargrave

Drive between Clifton and Coalcliff (Fig. 1.4). This section of Lawrence Hargrave

Drive has a history of slippage, rock falls and mud-slips. The area is geologically

unstable and is particularly vulnerable during periods of heavy rain.

During 1988 the road was closed between Clifton and Coalcliff on 30 April following

major rock falls, mud-slides and subsidence. The restoration work was carried out by

the Road and Traffic Authority (RTA) at a cost of $4.8 M and the road re-opened on

15 November. The majority of restoration work was in the Coalcliff area but some

repairs were also carried out on the "Clifton Fault" where further problems have now

been encountered.

1.5 STUDY METHODS


The present study has concentrated on geological factors affecting slope stability and a

literature review of general problems of, and strategies for, assessing slope stability.

Field work has concentrated on the mapping and investigation of landslides and rock

failures with the particular aim of gaining an understanding of the engineering geologica

features of the area between Clifton and Coalcliff. This study consists of an assessment

of the following factors.

(1) Previous studies of slope stability and engineering properties of rock sequences in

the Scarborough-Stanwell Park area.

(2) Geology of the upper Illawarra Coal Measures and lower Narrabeen Group in the

Scarborough-Stanwell Park area. This includes outcrop characterisation, the

thickness of sedimentary units, types of lithology, types of bedding, bedding

sequences and sedimentary structures.

(3) Structural geology in the slip area including the nature, size and orientation of

faults and faults planes; and the relationship between fault planes, weathering and

geomorphology; the nature, spacing type, regularity and orientation of joints in the
9

sandstone and shale units. The relationship between joints and orientations of cliff

faces and rates of erosion of stratigraphic units in different type of exposures is

also considered.

(4) Permeability of the various rock units for predicting groundwater movement based

on observation of surface seepage. The relationship between predicted groundwater

m o v e m e n t and the structural and stratigraphic features was also considered.

(5) Petrology of coarse- and fine-grained sandstone and the intervening shales in fresh

drill hole samples. Petrology of the same units in outcrop samples to determine

the effects of surface weathering. Petrology of talus material, with special

reference to material in slip zones. Scanning Electron Microscope ( S E M ) studies

of clays and cements occurring in joints and on slip planes have been carried

out.

(6) Detection of slip surfaces, water table level and lithological boundaries for

assessing the possible effects of fault activity and earthquake shaking.

(7) R o c k strength and durability - comparison of fresh, weathered and talus material.

The relationship between engineering test results and penological characteristics

of each unit.

(8) Soil instability, including case examples of slips in the area.

(9) Rock instability in the Narrabeen Group, including case example of rockfall and

toppling in the area.

(10) Predicted rates of weathering based on petrography.

(11) Predicted controls on land stability - fall and slip.

(12) Preventative measures that can be used to minimise rockfalls, toppling and slips

affecting the road between Clifton and Stanwell Park.

(13) preventive measures that can be used to minimise slippage of talus deposits in the

area.
10
1.6 MASS MOVEMENT
The study of mass movement, especially landslides, encompasses over one hundred years

of specialist work. During that time understanding the form and processes governing

these displacements of material has increased considerably. As the gaps in our

knowledge are filled, so the variety and complexity of moving soils and rocks becomes

more obvious.

The term mass movement is used here as a general term to include falls, topples, slides,

flows and or slumps along distinct slide planes or zones of sliding. Gravity is the

principal driving force; the movement is directed down and out, and the displaced

material may include soil (regolith), bedrock and or artificial fills. The term, as used

here, includes rockfalls, topples and debris flows which involve little or no true slidi

on a slide surface (Hansen, 1984).

Consideration of soil creep, which occurs without a well defined failure surface, is

excluded from this thesis. Creep is taken to refer to mass movements at rate of less

than about 0.06 m/y (Varnes, 1978). Many landslides creep before readily observed

movement occurs.

1.6.1 LANDSLIDE TERMINOLOGY


The terminology used to describe landslides is defined in Figure 1.7. The definitions

largely follow those of Varnes (1978).

1.6.2 TYPE OF MASS MOVEMENT


Landslide classification has been based on:

(1) the shape of the slide surface (landslide morphology);

(2) the mode of movement;

(3) the rate of movement;


11

(4) the type of material involved;

(5) the age of material;

(6) the age of failure; and

(7) various combinations of the above.

M a n y classifications of landslides and discussions of mechanisms of failure have been

reported in the literature, e.g. Sharp (1938), Terzaghi (1950), Skempton (1953, 1964),

Varnes (1958, 1978), Selby (1967), Hutchinson (1968, 1988), Skempton and Hutchinson

(1969), Zaruba and Mencel (1969), Blong (1973), East (1978), Chowdhury (1980),

Crozier (1986) and Vaunat et al. (1994).

This section considers aspects of only a few classification schemes. In a computer-

based study of landslide morphology, Blong (1973) concluded with a very simple and

useful classification whereby the primary divisions recognised were slide, rotational slide,

flow and fall (Fig. 1.8).

The simplest morphometric classification of slope failure is based on the D/L ratio

defined by Skempton (1953) where D is the maximum thickness of the landslide and

L is the m a x i m u m length in the direction of maximum slope (Fig. 1.7). Values of the

ratio increase from flows, through slides to slumps as shown in Table 1.1. The ratio

is useful, but it is difficult to use with accuracy where the failed mass has been

truncated by stream or other action.

Table 1.1 Typical D / L % Ratios for various landslide types based on data in

Skempton (1953), Selby (1967) and East (1978).

Landslide type D/L %

Flows 0.5-3.0

Slides 5-10

Slumps 15-30

3 0009 03155309 7
12

Chowdhury (1980) classified slides according to their causes:

(1) landslides arising from exceptional causes such as earthquakes, exceptionally high

precipitation, severe flooding, accelerated erosion from wave action, and

liquefaction;

(2) ordinary landslides, or landslides resulting from known or usual causes which can

be explained by traditional theories; and

(3) landslides which occur without any apparent cause. Although all of the above

classification schemes are simple and useful they are not widely used, and it is

often necessary to refer descriptions of slope failures to a more detailed

classification.

Varnes' classification (1978) is very useful for many research relationships on slope

stability. The basic types of slope movement in the classification are summarised in

Figure 1.9. An important reason for the acceptability of Varnes' classification was the

presentation of a set of three-dimensional diagrams such as those reproduced in Figure

1.10. Varnes recognised six principle types of mass movement:

(1) Falls

Falls are slope movements on steep slopes where a discrete mass of material, regardless

of size, is detached and moves downslope by travelling through the air, bouncing, or

rolling. More detailed classifications of falls are based on the type of material. Rock

falls involve bedrock, while debris falls involve coarse-grained fragments, and earth

falls involve fine-grained aggregates of material.

(2) Topples

Topples occur when a tensile failure in the rock mass causes it to rotate about a point

below its centre of gravity. These types of movement usually occur on steep slopes and

may terminate as a fall or slide, depending on the geometry of the slope below the

point of rotation. Topples can occur in any cohesive material and may range in size
13

from very small to extremely large. The size of the mass that topples is controlled by

the substance and mass characteristics of the topple material.

(3) Slides

Slide m o v e m e n t is either rotational or translational. Rotational slides occur in earth

materials w h e n the strength of the slide material is nearly equal to the strength along

discontinuities in the rock or soil mass. Rotational slides commonly appear as slumps

of material along slopes, road cuts and fills. Translational slides occur along planar or

gently undulating failure surfaces. These types of slides usually occur in earth materials

where the strength of the slide material is greater than the shear strength along

discontinuities in the rock or soil mass.

(4) Spreads

In spreads the sense of movement is nearly horizontal, and the earth material fails both

by shear along a failure surface, and by tension or extension along one or more nearly

vertical surfaces. This type of movement requires that some underlying geologic unit

fails and moves outward, carrying the overlying materials. The stability of the slope

is controlled by the geologic units at the site and by the loading conditions.

(5) Flows

Flows are slope movements in which the mechanical properties of the slope material

behave as a plastic body, viscous fluid or true fluid. In bedrock, these include spatially

continuous deformations, and deep creep involving extremely slow and generally non-

accelerating differential movements along relatively intact units. In soil, movement

occurs within a displaced mass, whereby the form or apparent distribution of velocities

resemble that of a viscous fluid.

(6) C o m p l e x slides

Landslides m a y exhibit a combination of two or more of the five principal types of

m o v e m e n t listed above.
14

Varnes (1978) also defined another useful term: multiple movements are those in which

repeated failure of the same style occur one after the other (Fig. l.lOe).

1.6.3 FACTORS CAUSING MASS MOVEMENTS


It is important to recognise the conditions that cause a slope to become unstable and

the factors that trigger the movement. An early recognition of some of the

environmental influences on slope stability was made by Terzaghi (1950) who listed

external changes which increased shearing stress, and internal changes which decreased

shearing resistance. Rainfall and earthquakes were recognised as contributing to slope

failure. A modified version of Terzaghi's work is presented in Table 1.2.

Zaruba and Mencel (1969) elaborated upon this as follows: "preventive treatment of a

landslide or an area susceptible to sliding must be based on a detailed, integrated

geological investigation. It is necessary to study the geological structure of the area,

the petrographical and physical properties of rocks, and the hydrogeological conditions.

As the form of a slope is the end product of geological processes of the past, the

morphological history of the slope must also be understood". These statements certainly

apply to the Illawarra area where, as indicated previously, marginally stable or unstable

colluvial masses exist on many slopes.

General procedures for engineering geological investigations of slopes have been outlined

in numerous references and will not be reviewed here. Procedures described by

Dearman and Fookes (1974), Deere and Patton (1971), Patton and Hendron (1974) and

Bhandari and Thayalan (1994) are particularly applicable to the Illawarra area. This

section attempts to explain the relationships between landslide occurrence and

environmental and geological factors.

(1) Relationship between slope movement and precipitation

Rainfall is generally accepted as one of the chief factors controlling the frequency of

landslides. The magnitude of its influence depends on climatic conditions (such as the
15

distribution of precipitation and changes in temperatures), topography of the area, the

geological structure of slopes, and the permeability and other properties of rocks and

soils.

(2) Slope

Usually, landsliding will occur more readily on steep slopes. While reality is more

complicated than this, it is often possible to determine lower limits (thresholds) of slope

below which landslides are unlikely to occur. But care must be taken in transfering

information about threshold slopes even to an adjacent area. For example, Dunkerley's

(1976) work at Razorback, south of Sydney, indicates that for this area of Wianamatta

Shale with numerous landslides, the threshold angle for earthflows is 11 degrees. While

at West Pennant Hills where the parent materials are also Wianamatta Shales, landslides

occur on slopes as gentle as 6 degrees (Fell, 1985).

(3) Slope shape

Slope shape both across the slope and down the slope m a y affect landslide location.

Slope shape in a downslope direction m a y be a reflection of slope steepness. O n e study

that examined landslide-slope shape relationships (Waltz, 1971) indicated for the San

Fransisco B a y area that landslides were commonly located on sites that are relatively

convex in both downslope and across slope direction.

(4) Stratigraphy

The stratigraphy at a particular site has a major influence on the slope stability.

Therefore, one of the first steps in the investigation of a slope should be the

determination of the stratigraphy of the area. Details of stratigraphy should be

determined to a degree commensurate with engineering requirements of the region.

In stratigraphic studies, particular attention should be directed toward recognition and

description of thick sequences of weak rocks such as claystones, thin marker beds such

as coal and clay seams, carbonaceous shales and old failure surfaces or shear zones.
16

Coal seams require special emphasis. They commonly act as aquifers and their

underclays often contribute to failure surface development in colluvial slopes.

(5) Structural geologic factors

Two main groups of geologic factors distinguish slope stability problems in soil from

those of rock. One group of geologic factors is related to the structural defects found

in rock masses and the special strength problems that result, whereas the other group

is related to special groundwater conditions which are more commonly associated with

rock masses than with soil. The critical groundwater conditions are often a direct

consequence of the presence of structural defects. In general, rock masses are best

considered as possessing anisotropic strength, permeability and deformability

characteristics to a much greater and more significant degree than soils.

The presence, continuity, spacing, orientation and nature of joints and bedding in the

weathered rock beneath the soil will in many cases be the dominant control over

landsliding. Shear zones and faults can also have a major effect on hydrological

conditions since they act as aquifers or aquitards depending on particular circumstances.

A search for faults or shear zones having low shear strengths due to previous

displacements is very important. The search is aided by the knowledge that faults or

shear zones are characteristically associated with particular geologic environments.

These consist of: (a) joints and faults subparallel to, or in secondary or conjugate

alignment to, regional faults; and (b) bedding plane faults and joints in shales where

they are interbedded with other rock types. Item (b) above is particularly common in

folded or inclined sediments adjacent to thick layers of a relatively less deformable rock

such as sandstone.

(6) Hydrogeological factors

Water pressure within a rock mass acts perpendicular to the surfaces of the

discontinuities. When there are many joint sets with different orientations and when the

joint spacing is small, the water pressure within the rock mass can be treated in a
17

similar model to that used for soil slopes. However, when the distribution of joint

orientations is anisotropic, and when the spacing between joints is increased, m a n y

unusual distributions of water pressure can result. In rock masses it is possible to have

the water pressure, and hence the corresponding disturbing force, change appreciably

from one joint to the next. A s shown in (Fig. 1.11) the water level is m u c h lower in

joint a-a than in joint b-b. A s a result, the magnitude of the force P b due to the

hydrostatic pressure along joint b-b is several times the force P a acting normal to joint

a-a. Groundwater fluctuations affect slope stability in both rock and soil slopes.

Groundwater levels are likely to fluctuate much more in rock slopes than in m a n y soil

slopes due to the smaller percentage of void space in rocks and the more open joint

systems. The effects on the groundwater table of a 2.5 c m rainfall which entirely

infiltrates into a porous soil slope and a low porosity rock slope is shown in Figure

1.12. In porous soils (Fig. 1.12a) 2.5 c m of rainfall can produce an 8 c m to 25 c m rise

in the groundwater level assuming porosities of 3 3 % to 1 0 % , respectively. The same

rainfall on a rock slope could produce increases in groundwater levels of the order of

m a n y metres (Fig. 1.12b).

Fortunately, the rock mass adjacent to many rock slopes becomes more permeable

because the joints open due to stress relief. This zone of more open jointing serves to

retard the development of high water pressures near the slope surface.

(7) Earthquakes

Earthquakes affect slope stability as follows, (a) Earthquakes produce horizontal and

vertical accelerations in soil masses. The horizontal acceleration m a y reach 0.5 g

(gravitational acceleration) or more, altering the distribution of forces in hillslopes in a

manner equivalent to a temporary steepening of the slope, (b) Earthquakes can change

the magnitude and distribution of the pore water pressure and increased pore water

pressure reduces soil shear strength. Rapid increase of pore water pressure in some
18

coarse-grained soils occurs by repeated shear stress fluctuations. In loose sandy soils

this cyclic shear loading may lead to liquefaction, i.e. total loss of shear strength.

Keefer (1984) suggested that rockfalls, rockslides, soil-falls and soil slides can be

triggered by the weakest seismic activity while deep seated slumps and earthflows are

generally initiated by stronger and perhaps longer ground shaking.

1.7 MAIN CAUSES OF LANDSLIDES IN THE ILLAWARRA AREA


Observation and study of areas of instability in the Illawarra region have shown that

many slides are associated with the Illawarra Coal Measures and Narrabeen Group.

Different thicknesses of talus cover the slopes over a large area. Depending on their

topographical position, these talus deposits may be derived from the Hawkesbury

Sandstone, Narrabeen Group, Illawarra Coal Measures or combinations of these older

strata.

Landslides have taken place in the Illawarra area due to:

(1) Properties of the talus

The talus usually consists of sandstone boulders in an iron-stained clayey matrix. With

any heavy rain, pore water pressure rises, decreasing the shear strength of the talus. As

the shear strength reduces to the level of the applied shear stress, the talus matrix is

mobilised, resulting in mass movement. Where natural drainage occurs, the talus has

become consolidated and, with compaction and deposition of cementing materials from

solutions, it is relatively impervious to percolating waters. Such talus is often stable on

gentle topographic slopes.

(2) Rock jointing and erosion by the sea

The properties of the underlying rocks have important effects on the nature and stability

of talus. Sandstone cliffs along the sea are vertically jointed and blocks of rock will

break off leaving vertical faces. Rockfalls occur in these places because the toe of the
19

slope is eroded by the sea; relaxation of the material above produces toppling along

joint faces.

Soft rocks and weathered volcanic sandstone (Illawarra Coal Measures) are exposed at

sea level and at the base of some landslides; slips occur from fretting and weathering

of these lower strength rocks.

(3) Properties of the Illawarra Coal Measures

The Illawarra Coal Measures consist of a repetitious sequence of sandstone interbedded

with shale and coal. Coal beds commonly act as aquifers, with shale acting as

aquitards. Landsliding can be related to the presence of the aquifers.

(4) Petrology of Illawarra Coal Measures and Narrabeen G r o u p

The volcanic detritus is present in both the sandstone and shale units either in form of

detrital grains of volcanic rock or as fine volcanic ash. During post-depositional

alteration and diagenesis the original volcanic glass in the ash and matrix of larger

grains has devitrified to produce smectite clays. These clays not only cause swelling

and shrinkage near the surface as a response to wetting and drying, but also reduce the

permeability of the near the surface rock mass. This latter factor increase the aqueous

pore pressures and hence increase the likelihood of surficial mass movement of both the

rock mass and the adjacent talus deposits.

(5) Geotechnical Properties of the Illawarra Coal Measures and Narrabeen G r o u p

High proportions of expansive clay minerals were detected in volcanic rock fragments

(cherts) which suggest that clay softening in the presence of water is important in

controlling moisture related reduction in strength of sandstone in the Illawarra area. The

nature of the cement (kaolinite) and the rate of weathering also influences slake

durability. Weathered shale and Claystone in the Narrabeen Group have low to very

low durability. Weathered sandstone in the Narrabeen Group is moderately strong to

strong whereas measured values from weathered shale and claystone are moderately

weak to moderately strong.


20

(6) Local structural geology

The relationship of landslides to structural geology in this area appears to be very

important. Local structural features such as fractures and faults influence the

underground water circulation. They cause an increase in local water flow and appear

to be directly related to landslides. Regional tilt is about 5 to the north-northwest

causes joint blocks initially tilt landwards. The intersect of joints and bedding planes

causes additional surfaces of weakness which accompany to lithological changes

(sandstone to shale or sandstone to claystone and sandstone interbedded with claystone

and coal causes slope instability in the Illawarra area.

(7) Effects of long-term weathering

M a n y rock falls and topples occur along the Lawrence Hargrave Drive between Clifton

and Coalcliff where the bedding is nearly horizontal and jointing is approximately

vertical. Most effects of weathering are concentrated along the cliffs and in the soft

interbedded shale units. Fretting and weathering of shales have undermined sandstone

layers. This causes large blocks of the overlying sandstone to break off and

considerable tonnages of rock have fallen on m a n y occasions, often completely blocking

the road.

(8) Influence of water

Increase in the pore water pressure within a talus slope affects its stability. Lack of

adequate drainage can cause a rapid increase in the pore water pressure after heavy
i

rainfall; this leads a decrease in the shearing resistance of the talus material, and causes

mass movement.

Water flow within the rock mass is concentrated along discontinuities at the base of

sandstone units, i.e. between claystone and sandstone beds. This increases the rate of

weathering of the claystone and causes fretting and weathering of the sandstone as well

as toppling failures in the area.


21

T o understanding and quantify the causes of failure and to provide the basic data

necessary for later study, it was first necessary to study of regional geology and carry

out laboratory testing of rock and soil material involved. Fresh and weathered

sandstone, shale, claystone and talus with varying moisture content were sampled over

the area under investigation and tested for a range of laboratory properties. Particular

emphasis was placed on understanding the change of the petrological and mechanical

properties with an increase in weathering.


22
23

CHAPTER 2

GEOLOGICAL AND GEOGRAPHICAL REVIEW OF THE

ILLAWARRA REGION

2.1 REGIONAL GEOLOGY


The Sydney region forms part of an interconnected network of Permo-Triassic basins

in eastern Australia. It extends from the highly deformed middle Palaeozoic Lachlan

Fold in the west to the continental margin in the east. Its eastern extent was

terminated at the outer edge of the present continental shelf by the opening of the

Tasman Sea (Veevers et al, 1991) and the basin is bordered by the New England

Fold Belt in the northeast (Fig. 2.1).

The rocks of the Sydney region are dominantly of sedimentary origin and have been

deposited within a broad zone of subsidence known as the Sydney Basin (Fig. 2.1).

-Deposition took place from carboniferous to the latter part of the Triassic upon a

basement of Early to Middle Palaeozoic metamorphic and igneous rocks-(Rickwood,

1985). For the most part, the strata are conformable and close to horizontal.

The present Sydney Basin succession comprises up to 6 km of Permo-Triassic

sediments (Mayne et al., 1974). The thickest part of the succession is adjacent to

the New England Orogen in the Newcastle area but the sequence is also thought to

thicken eastwards towards the continental shelf. The succession gradually thins to

the south and west. In the southern Sydney Basin, where the present study was

conducted, the maximum thickness is more than 2.5 km. The depositional and

tectonic history of the basin has been well described in the literature (Conolly and

Ferm, 1971; Mayne et al, 197'4; Herbert, 1980a; and Branagan 1985). A marine

transgression during the Early Permian allowed the deposition of thick sequences of
24

marine sediments (Shoalhaven Group) and a major regression in the Late Permian

resulted in the deposition of the Illawarra Coal Measures about 250-260 million years

ago.
Deposition in the Late Permian and Early Triassic took place essentially in marginal

marine to fluvial environments with numerous coal swamps. Deposition is postulated

to have extended into Jurassic times with an hiatus in the Late Triassic (Herbert,

1980a).
Permian deformation mainly produced broad folding and some faulting in the

northern part of the basin near the thrust margin with the N e w England Fold Belt.

The basin contains minor occurrences of irregularly distributed Cainozoic deposits.

Occasional volcanic activity in the form of dykes and other intrusive bodies persisted

from Permian to Tertiary.

2.2 REGIONAL STRUCTURE


The Sydney Basin developed its structural entity in the mid Permian, after the N e w

England Orogeny (Herbert, 1980a). Subsidence of the basin started with initial

sedimentation in the Newcastle area to the north of the present study area, and

comprised molasse sediments derived from the N e w England Fold Belt (Herbert,

1980a; Roberts and Engle, 1987). Subsidence continued throughout the Early

Permian and transgressive shallow marine and paralic sediments were deposited

(Herbert, 1980a; Roberts and Engle, 1987). Subsequently, during the Late Permian,

intense crustal movement occurred and N e w England Orogen was uplifted to the

north of Hunter Thrust System during the Hunter Orogeny (Scheibner, 1976).

The Hunter Orogeny was immediately followed by molasse sedimentation and marine

regression. Most of major coal deposition in the Sydney Basin was in alluvial and

deltaic environments during this Late Permian regressive phase (Bamberry, 1992).
25

Towards the end of the Late Permian, coal measure sedimentation terminated and

was followed by the deposition of predominantly fluvial sequences, in the Triassic.

The time at which sedimentation ceased in the Sydney Basin is not known.

Palynological studies of spores preserved in diatremes in the Sydney Basin indicated

that they are of Early Jurassic age (Crawford et al, 1980). Therefore, it has been

postulated that deposition in the basin continued at least up to Early Jurassic with a

hiatus in the Late Triassic (Herbert, 1980a). Recent investigation by Jenkins et al.

(1993) on the continental slope off the southeastern coast of Australia suggest that

some deposition may have also occurred in the Cretaceous.

-Studies have indicated that the present continental margin along the southeastern

coast of Australia developed as a result of continental break-up in the Late

Cretaceous. Palaeomagnetic, radiometric and fission track data indicate that rifting

began approximately 100 Ma and sea floor spreading (drifting) occurred

approximately 85 Ma (Weissel and Hayes, 1977; Shaw, 1978; Jones and Veevers,

1983; Moore et al, 1986; Dumitru et al, 1991; Veevers et al, 1991). The major

uplift and erosion of Sydney Basin sequence is believed to have been initiated in

relation to this spreading event in the Late Cretaceous-(Mayne et al, 1974; Oilier,

1982; Moore et al, 1986).

The Sydney region does not show evidence of strong deformation but detailed

investigations of the rocks shows that gentle deformation occurred both during and

after the main period of deposition in Permian and Triassic times.

In general, a variety of structural features can be recognised in the region. -These

features consist of broad depressions and gently inclined plateaux, folds, warps (Fig.

2.2), fault zones, faults and joints. Folds are gentle with north-northeast to northeast

axial trends in the central and northern parts of the basin (Branagan et al, 1988).

Fault zones have a patchy distribution. Where faulting occurs it is common to find

a variety of faults (normal, reverse, low angle thrusts and strike slip) within the area.
26

Normal faults occur in a number of orientations but faults trending northwest, n

northeast and northeast appear to be the most common (Norman and Branagan, 1984

Branagan, 1985). Thrust faults in the Sydney region have orientations ranging fr

northwest to north-northeast and northeast, with occasional more east-west orie

thrusts. Most are low angle or bedding-parallel structures (Branagan et al, 1988)

The Sydney region is also characterised by sub-vertical north-northeast trendin

zones which parallel an important joint direction (Norman and Branagan, 1984). T

dominant movement appears to be strike-slip, although some normal and reverse

displacements may also be present.

Jointing is widespread and at least four main trends have been recorded in the

(Fig. 2.3). These joints show a wide variety in their shape, continuity, inclina

and openness. Vertical north-northeast planar joints are continuous for many met

throughout the sandstone units and control the orientation of many cliff-lines

stream courses. Many joints are the passageways for groundwater and this results

a variety of rock conditions. Some joints and adjacent rock masses are strongly

leached and usually weakened while elsewhere significant deposition of iron oxi

cements has occurred often producing a more resistant material.

Jointing in the shale is tighter than in the sandstone when first exposed. Thes

joints maybe coated by calcium carbonate, clay or pyrite. Open joints in sandsto

maybe coated with these substances and with quartz or iron oxides. Joint faces i

sandstone retain their character, whereas joints faces in shale become rapidly

modified and exfoliate on exposure.

2.3 ILLAWARRA AREA

2.3.1 GEOMORPHOLOGY

The niawarra area is located in the southern part of Sydney Basin and comprises

tableland, the coastal plain, and the escarpment and slopes (Figs 1.4 and 1.5).
27

tableland was named the Woronora Plateau by Branagan and Packham (1967). It

ranges in elevation from about 360 m in the north to a m a x i m u m of 760 m in the

south. Hawkesbury Sandstone crops out over m u c h of the plateau, ranging in height

about 350 m behind Stanwell Park to 469 m in the M t Keira area.

The niawarra coastal plain stretches southward from Coledale where softer rocks,

particularly of the Illawarra Coal Measures, are exposed at or above sea level. It is

widest in the south due to the presence of large streams and more rapid erosion of

the weathered marine rocks. The Illawarra coastal plain has been formed by

westward recession of the plateau giving a faceted slope or scarp (Fig. 1.5). The

Hawkesbury Sandstone forms prominent cliffs on the crest of the escarpment.

Differential erosion of the underlying interbedded sandstone and claystone has

produced structural benches on the escarpment, with steep sandstone rises and

relatively flat claystone slopes. Debris or talus partly covers the structural benches

and bluffs and results from erosion of the interbedded sandstone and claystone of the

area. The Illawarra Coal Measures do not form major benches because the sandstone

units are thinner and less resistant to erosion.

The escarpment bounds the area along the western edge; and hence the coastal

lowland varies in width, from 5 to 20 k m (Fig. 1.4). The northern coastline is

characterised by steep cliffs of sandstone; these cliffs range in height from 3 m to 70

m , accompanied by alternating bays and barrier beaches of variable width. The

exposures along the cliffs are generally good, as a result of wave abrasion, whereas

little exposure is found on the cliff tops which have been covered by surficial

deposits, especially in residential areas.

Rock platforms with cliff notches are extensively developed in the softer Permian

sequences. Harder Triassic rocks north of Clifton form narrower platforms. The

platforms are composed of sandstone, which is more resistant to erosion by the sea
28

than the interbedded shale. On these rock platforms, systems of joints are

particularly well exposed, especially between Clifton and Coalcliff.

2.3.2 CLIMATE

In general, the Illawarra area has a temperate marine climatic regime. It is a

characteristically moist climate with no major dry season. The rainfall is fairly

evenly distributed throughout the year although slightly more rain falls in late

summer and early winter. Factors such as topography, structure of the region and

the nature of prevailing air masses all affect the rate of rainfall in the area.

Higher rainfalls are recorded on the plateau than on the coastal plain. About 1600

m m per year falls on the high ground west of the escarpment. Approximately 1500

m m per year falls along most of the escarpment while around 1200 m m per year

falls on the coastal plain. The average annual rainfall contours for the study area are

shown (Fig. 2.4).

It is k n o w n that higher rainfalls occur at higher elevations where temperature and

evaporation are lower. In the Wollongong area precipitation exceeds evaporation for

three months per year. Near the crest of the escarpment, precipitation is more than

evaporation for all months. The annual average excess is about 700 m m and results

in an increase of soil moisture, which leads to mass movement along the niawarra

escarpment (Chowdhury and Young, 1987).

The important point is that intense storms accompanied by a high rainfall occur

within a short period of time in this area. Often slips were caused by these rainfalls

(e.g. the landslide in the Coalcliff area in April 1988).

Information from B o w m a n (1972) shows an average m a x i m u m temperature for the

hottest month, February, of 26.9C and an average m i n i m u m for coldest month, July,

of 8.4C with an average daily temperature over the whole year of 17.5C. Winter
29

winds are from the west and southwest while summer winds are commonly from the

south or northeast.

2.3.3 GEOLOGY

The study area is located in the southeastern portion of the Sydney Basin (Fig. 2.1).

It comprises a sedimentary sequence with several volcanic units. The rocks range in

age from Permian to Triassic. This part of the Sydney Basin has received

considerable attention from geologists because of its long history of coal mining.

Harper (1915) produced the first comprehensive description of the geology of the

Illawarra area. H e described the general stratigraphy and structure and recorded

many observations about the various coal seams. The work of Hanlon (1952, 1953,

1958) greatly expanded the knowledge of the area. H e revised and modernised the

stratigraphic nomenclature and described in considerable detail the stratigraphy of the

area between Coledale and Stanwell Park. M u c h of detail available about the

various rock sequences is due to Hanlon's work.

B o w m a n (1974) mapped the geology of Wollongong, Kiama and Robertson sheets at

a scale of 1:50000. Adamson (1974), in an investigation for Coalcliff Colliery,

added some detail not previously recorded. Chestnut (1981) mapped the

Wollongong-Port Hacking area with emphasis on engineering and environmental

geological hazards. Finally the Geological Survey of N e w South Wales published a

report on the geology of Wollongong and Port-Hacking (Sherwin and Holmes, 1986).

Although the geology of the niawarra has been studied for a century or more, few

studies have attempted to understand the slope stability in this area.

Slope instability in the northern niawarra almost totally occurs within the upper

Illawarra Coal Measures and Narrabeen Group. Therefore this study is concerned

with upper Illawarra Coal Measures, Narrabeen Group and associated talus. Other

formations are described only briefly.


30

2.3.4 STRATIGRAPHY

2.3.4.1 SHOALHAVEN GROUP

Shoalhaven Group in the southern Sydney Basin consists predominantly of alternating

sandstone and siltstone sequence (Jones, 1990). The Broughton Formation at the top

of Shoalhaven Group is the oldest rock exposed in the study area (Carr, 1983). It

crops out along the coastal plain in the Wollongong area. The sandstone is a red to

green-grey lithic sandstone with thin interbeds of siltstone and conglomerate. Five

tabular latite bodies are interbedded in the Broughton Formation in the Kiama area to

form the Gerringong Volcanic facies. One of them is the Dapto Latite Member

which occurs around Lake Illawarra. This volcanic unit is responsible for the higher

topographic area between Wollongong and Lake Illawarra. Its presence has

significantly influenced the evolution of the coastal plain in this region. The

maximum thickness of Broughton Formation reaches 370 m at Saddleback Mountain

(Bowman, 1980).

2.3.4.2 ILLAWARRA COAL MEASURES

The Illawarra Coal Measures are generally located at the base of the escarpment in

the northern Illawarra. The coastal plain is formed partly on this formation, although

Quaternary alluvium frequently overlies the coal measures. The idealised stratigraphy

of Illawarra Coal Measures is presented in Figure 2.5. This formation had been

studied by numerous authors, e.g. Harper (1915), Hanlon (1952, 1953, 1958),

Bowman (1974), Odins et al (1990) and Bamberry (1992).

In general the coal measures are subdivided into the Cumberland Subgroup and the

overlying Sydney Subgroup. The Pheasants Nest Formation is the basal formation

within the Cumberland Subgroup and hence the niawarra Coal Measures. It

represents a transition from the underlying marine Shoalhaven Group strata to a


31

fluvio-deltaic depositional regime. This unit consists of sandstone with shale,

siltstone and conglomerate interbeds. Two small lenticular coal seams are located

within the Pheasants Nest Formation. The overlying marginate marine Erins Vale

Formation is a fine- to medium-grained lithic sandstone with few shale and

conglomerate lenses.

The Sydney Subgroup disconformably overlies the Cumberland Subgroup and is

composed of numerous sedimentary formations, as well as many members. The

subgroup consists of interbedded quartz-lithic sandstone, grey laminated mudstone,

and nine coal units. The uppermost formation is the Bulli Coal which is the main

commercial coal seam in the niawarra area. It varies in thickness from 4 m in the

north to 1.5 m near Port Kembla and contains few claystone bands. Large areas of

the Bulli Coal have now been mined.

Sandstone units within the coal measures are commonly moderately to well sorted,

with fine- to medium-grained subrounded to rounded grains. Cement is usually

calcite and porosity is generally low (Table 2.1).

2.3.4.3 NARRABEEN GROUP

The Narrabeen Group is known to occur throughout the Sydney Basin. It extends

along the Illawarra coastal escarpment and also crops out to the west of the

escarpment. This group includes the main sequence of rocks along the coastal cliffs

between Stanwell Park and Scarborough, where it is particularly well exposed.

The lowest units of the Narrabeen Group are Late Permian and the upper unit is

Middle to Late Triassic in age. The thickness of the Narrabeen Group decreases to

the south. For example it has a thickness of 330 m north of Otford (Loughnan,

1963) while at Clifton it is 253 m thick (Hanlon, 1953).

The Narrabeen Group includes the Coal Cliff Sandstone, Wombarra Shale, Otford

Sandstone Member, Scarborough Sandstone, Stanwell Park Claystone, Bulgo


32

Sandstone, Bald Hill Claystone, Garie Formation and Newport Formation. The

Hawkesbury Sandstone overlies the Narrabeen Group.

2.3.4.4 COAL CLIFF SANDSTONE


The Coal Cliff Sandstone is the basal unit of the Narrabeen Group and overlies the

Illawarra Coal Measures. The thickness of the unit ranges between 6 and 20 m ,

being approximately 9 m thick in the type section at Clifton (Hanlon, 1953). The

Coal Cliff Sandstone is a light-grey, fine- to medium-grained, quartz-lithic and lithic

sandstone with a number of pebble and shale bands. It crops out in the coastal

section near Clifton and passes below sea level north of Coalcliff. Angular siderite

fragments up to 10 c m in size are c o m m o n in the basal Coal Cliff Sandstone.

This unit forms the roof of some colliery workings and is exposed underground for

several kilometres to the west of the Illawarra escarpment. In the some places

colliery roofs are less stable because the fine sandstone near the base of the Coal

Cliff Sandstone sometimes grades into shale.

2.3.4.5 WOMBARRA SHALE

The Coal Cliff Sandstone is overlain by 6 to 30 m of greenish-grey shale with lithic

sandstone interbeds. It is well exposed in road cuttings and cliffs south of Coalcliff.

A measured section of the Wombarra Shale (Fig. 2.6) illustrates the frequency of the

interbedded sandstone. The sandstone interbeds are generally quite thin, lenticular,

fine-grained and carbonate-cemented. Asymmetrical ripple marks occur at the base

of the Wombarra Shale at Coalcliff (Bowman, 1974).

Towards the top of the formation a thicker sandstone unit is called the Otford

Sandstone M e m b e r (Hanlon, 1952). This member is approximately 7 m thick but

varies because of lateral changes to shale.


33

A clear facies change occurs in the Wombarra Shale between Clifton and

Helensburgh (Loughnan, 1963). It comprises 7 7 % lutite at Clifton but contains only

4 0 % at Helensburgh. The lateral facies changes essentially affect its stability. If the

arenite content increases a little, weathering will decrease within the unit and stability

will be maintained.

2.3.4.6 SCARBOROUGH SANDSTONE


The Scarborough Sandstone rests on the Wombarra Shale. This unit is a massive

sandstone nearly 24 m thick for most of its outcrop length. C o m m o n l y the

Scarborough Sandstone is conglomeratic with coloured chert clasts especially in the

basal half. It consists of beds up to several metres in thickness which become finer

upwards. This unit comprises lithic to quartz-lithic sandstone with pebbles and

minor amounts of grey shale. The coarse nature of the sandstone has resulted in the

development of important cliffs between Stanwell Park and Clifton.

2.3.4.7 STANWELL PARK CLAYSTONE


This unit separates the Scarborough Sandstone from the Bulgo Sandstone. The

Stanwell Park Claystone is about 37 m thick at the type section between Clifton and

Coalcliff. It consists of interbedded green to chocolate shale and sandstone. Three

claystone intervals and two sandstone beds can be recognised.

The lower section of the unit consists of greenish-grey claystone and sandstone

which slowly changes upward into red-brown claystone and clay. The sandstone

beds are composed of weathered lithic fragments and are usually light greenish-grey

in colour. The relative proportion of claystone and sandstone varies but overall they

are sub-equal ( B o w m a n , 1974).


34

2.3.4.8 BULGO SANDSTONE

The Bulgo Sandstone, which rests on the Stanwell Park Claystone is the thickest unit

of the Narrabeen Group on the Illawarra coast. It forms prominent outcrops in the

area and between Coalcliff and Clifton. The Bulgo Sandstone is 120 m thick at

Clifton forming over half the total Narrabeen Group (Ward, 1980). It consists of

thickly bedded sandstone with intercalated siltstone and claystone beds up to 3 m

thick. Conglomerate is also present, especially toward the base.

A complete section of the Bulgo Sandstone is exposed in the cliffs south of Otford,

but it has not been studied due to the difficulty of access. Ward (1971a) subdivided

the formation into three facies: the basal "pebbly facies", the middle "volcanic facies"

and the upper "shaley facies".

The "pebbly facies" consists of pebbly sandstone like the Scarborough Sandstone.

This sandstone is mostly lithic. The "volcanic facies" consists of green sandstone

and shale. The sandstone consists predominantly of intermediate to basic volcanic

rock fragments which are altered to chlorite and iron oxides. The "shaley facies" has

a high proportion of siltstone. The shale is often red owing to hematite staining.

The Bulgo Sandstone has a higher proportion of quartz than of rock fragments.

Sandstone beds rarely exceed 4 m in thickness while the siltstone and shale interbeds

are usually less than 1 m thick. The red shale beds of the upper "shaley facies" are

up to 2 m in thickness.

2.3.4.9 BALD HILL CLAYSTONE

The Bald Hill claystone, which overlies the Bulgo Sandstone, crops out in the hills

near Otford and on the Mt Ousley road to the south. This formation is about 15 m

thick at its type locality in the Bald Hill area (Hanlon, 1953). It consists almost

entirely of claystone, but lithic sandstone interbeds are found towards the base of the

unit. Mottled chocolate and green claystone zones are common. The mineralogy of
35

the Bald Hill Claystone is quite simple. Kaolinite, hematite and/or siderite being

almost exclusively the only minerals present (Bowman, 1974).

2.3.4.10 GARIE FORMATION


Toward the top of the Bald Hill Claystone, thin beds of light coloured claystone

become more common. This upper zone passes into a mid-grey slightly

carbonaceous massive claystone, which is overlain in turn, by the Newport Formation

(Fig. 2.7). The Garie Formation is usually less than 3 m thick but it is a very good

marker horizon in the southern Sydney Basin.

2.3.4.11 NEWPORT FORMATION


The mid-grey shale and minor interbedded lithic sandstone of the Newport Formation

overlies the Garie Formation (Fig. 2.7). The formation is 18.5 m thick in its type

section on the coast 3 km north of Garie Beach (Hanlon, 1953), but is reduced to

11 m near Clifton.

Mud-rocks of this formation are thinly bedded. The dark-grey mud-rocks contain

plentiful plant fossils. Claystone beds consisting of sand-sized flakes of kaolinite,

with a large original porosity, are common in the Newport Formation (Bowman,

1974).

2.3.5 POST-NARRABEEN UNITS


2.3.5.1 HAWKESBURY SANDSTONE
This unit is a flat-lying Middle Triassic quartz sandstone with an areal extent of

about 20,000 km2 and a maximum thickness of 250 m (Standard, 1969; Conaghan,

1980). The Hawkesbury Sandstone crops out at the top of most the niawarra

escarpment. It forms a resistant plateau to the west of the escarpment, which gently

dips to the northwest. The formation has a thickness of about 180 m at Stanwell
36

Park. It contains a minor amount of mudstone, interbedded with fine sandstone, but

it consists dominantly of sandstone beds (Jones and Rust, 1983) typically 2 m to

5 m but up to 15 m in thickness. Transition into conglomerate is seen in some of

the sandstone beds. Strong cross-bedding is common in the Hawkesbury Sandstone

and it was discussed by Bowman (1974), Conaghan and Jones (1975), Conaghan

(1980) in some detail. The interbedded mudstone is very prone to weathering upon

exposure. The Hawkesbury Sandstone is often involved in rockfalls from the

escarpment,

2.3.5.2 WIANAMATTA GROUP


The Wianamatta Group crops out on the plateau to the west of the Elawarra

escarpment and overlies the Hawkesbury Sandstone. It consists of interbedded grey

shale, lithic and quartzose sandstone, and has a maximum thickness in the study area

of nearly 15 m. Helby (1973) has suggested a Middle Triassic age for the

Wianamatta Group.

2.3.5.3 IGNEOUS ROCKS


Illawarra area has numerous intrusive igneous bodies with one significant Tertiary

extrusive unit. The flow is called the Robertson Basalt. It covers the Robertson

plateau and is an alkaline basalt with a maximum thickness of 100 m (Sherwin and

Holmes, 1986).

A few small dykes and sills crop out, but they are usually weathered to clay.

Bowman (1974) indicated that most of the dykes have intruded along tension joints,

although Harper (1915) reported some intrusions along faults.

These igneous bodies locally affect the slope profile. They may typically weather to

montmoriUonite and are deleterious for stability.


37

2.3.5.4 T E R T I A R Y A N D Q U A T E R N A R Y DEPOSITS

Significant areas of the coastal plain and escarpment slope are covered by talus

which is Quaternary in age. It has a variable thickness below the Illawarra

escarpment. Generally the talus consists of large sandstone blocks, derived either

from the Hawkesbury Sandstone or from sandstone units in the Narrabeen Group, set

in a clayey matrix which is frequently iron-stained or leached.

This process is directly related to slope stability and is discussed more fully later.

2.3.6 STRUCTURAL GEOLOGY

The niawarra area lies near the southern edge of the Sydney Basin. The structural

geology of the area has been discussed by David (1896), Harper (1915), Wilson et

al (1958), Bowman (1974) and Mauger et al. (1983). In general, the strata in the

area have a regional dip of a few degrees to the north-northwest towards the centre

of the basin. The escarpment has a trend toward north-northeast. Anticlinal and

synclinal crenulations are developed on the eastern limb of the Camden Syncline

(Fig. 2.8).

2.3.6.1 FOLDS

A series of gentle folds occur along the southeastern edge of the Sydney Basin (Fig.

2.8). The axes trend about 155 and gently plunge towards the centre of basin.

Bowman (1974) suggested that the folding may have resulted from crustal

foreshorting during the deformation of the basin. Wilson et al, (1958) noted the

thickening of units in the synclines and thinning of units on the crest of anticlines

and he pointed out that sedimentation was contemporaneous with deformation.

Jakeman (1980) also discussed about the relationship between the formation structure

and thickness in Permo-Triassic succession of southern coalfield in the Sydney Basin.


38

According to Branagan and Pedram (1990) the monoclinal folding is associated with

faulting and may be associated with stress relief along a 105 -striking regional ac

(perpendicular to fold axis) joint set.

Several domes are developed in the area. Only the two largest, Mount Lindsay

Dome and Mount Burke dome are shown in Figure 2.8.

2.3.6.2 FAULTS

Most faulting in the area is normal faulting, although strike-slip and high-angle

reverse faults are known to exist. Wilson et al. (1958) recorded fault directions

consistent with the 110 and 155 folded directions. As well they reported another

set of faults striking approximately 050 with throws of less than 3 m. These faults

are generally difficult to recognise on the surface but are commonly encountered in

underground coal mines.

Generally they are downthrown to the north in the northern part of the niawarra area

(e.g Clifton Fault). Most of the major faults are clean breaks without a crush zone.

They usually form a series of small en echelon faults or fault zones (Fig. 2.9) and

horst and graben structures.

Bowman (1974) stated that three groups of faults occur (110, 155 and 005 in the

area (Fig. 2.10). The first two group correspond with those of Wilson et al (1958).

The lack of the 050 set in Bowman's data probably results from the difference

between underground and surface observations and hence measurements. Of these

faults directions, faults striking 110 are the most numerous. These directions

correspond to the tensional directions associated with the two directions of synclinal-

anticlinal folding. The displacement on the faults striking 005 is generally less than

5 m. Faults striking 110 and 155 can have much greater displacements. Minor

faults are normal faults with steep dips toward north. Most of the minor faults are

fracture zones several hundred metres across (Sherwin and Holmes, 1986). Evidence
39

exists that some faults m a y have been active during deposition, with a greater

thickness of sediment developing on the downthrown side and with a decreasing

throw up the section. For example, Figure 2.11 shows that the throw of the Jetty

Fault at Coalcliff decreases on ascending stratigraphically (Hanlon, 1953). Faulting

has taken place along the major joint direction and appears to be a tension feature.

Also a dyke occurs along the tension feature in this area.

2.3.6.3 JOINTS

Generally joints are the most significant structural feature in the area. Folding of the

sediments is commonly negligible and the strata are very close to horizontal.

Faulting is c o m m o n and fault zones influence the groundwater movement. Faults

usually play a similar role to jointing in assessing slope stability in this area.

Many workers have studied the distribution of joints in the Illawarra region,

including Dickson and Weber (1966), Connelly (1970), B o w m a n (1974) and Mauger

et al (1983). They found that there are four sets of prominent joint directions, at

005, 055, 105, 155. These directions are similar to the regional pattern (Fig. 2.3)

and are parallel the be (parallel to fold axis) and ac (perpendicular to folded axis)

directions of the two main fold orientations (Fig. 2.12). Therefore the joint sets are

probably tensional features resulting from stress relief after folding. But C o o k and

Johnson (1970) studied joint patterns in ironstone intraclasts and in the ironstone

layers from which they were derived on the coastal platform north of Wollongong.

They concluded that the jointing pattern present in the sediment, a set striking 078

and 105, was developed at the time of sedimentation. These suggests that the

regional northwest folds m a y have been synsedimentary - a concept confirmed by the

thickness trends for the Bulli Coal (Jakeman, 1980). Memarian (1993) also studied

the fracture history of the Coal Cliff Sandstone at Coalcliff. H e stated that these

joints are extensional in origin and formed from tectonic stresses during burial.
40

2.7 STRESS FIELDS


Many authors have commented on the stress fields affecting the Sydney Basin

including Scheibner (1976), Dolye et al. (1968), Denham (1980), Herbert (1980a),

Denham et al. (1981), Gray (1982), Everingham et al. (1987), Scheibner (1987),

Fredrich et al (1988), Greenhalgh et al. (1989), Michael-Leiba (1989) and Stone

(1990). Most of them agree that the Sydney Basin was subjected to east-west

compression from commencement of sedimentation in the Permian through most of

the Mesozoic. This east-west compression was probably related to subduction to the

east of the Australian continent (Scheibner, 1976).

In the Illawarra region the principal feature associated with this east-west stress fiel

is the Camden Syncline. In addition to this stress field, evidence exists that the

rocks were also subjected to northeast-southwest compression in much the same

period. This stress field caused the anticlinal-synclinal crenulations on the eastern

limb of the Camden Syncline.

Since the Cainozoic the basin has been subjected to north-south compression (Gray,

1976), features such as the monoclines and recent earthquakes are associated with

this later stress field.

Scheibner (1976) considered that this stress field was associated with movement of

the Australian plate away from the Antarctic plate.


41

CHAPTER 3
GEOLOGY OF THE UPPER COAL MEASURES AND LOWER NARRABEEN
GROUP IN THE SCARBOROUGH-STANWELL PARK AREA

3.1 INTRODUCTION
The lower Narrabeen Group and upper Illawarra Coal Measures in the Scarborough-

Stanwell Park area are essentially flat-lying strata consisting of repeated beds of

sandstone, shale, claystone and coal seams. The coal measure rocks consist of

tuffaceous sandstone, carbonaceous siltstone, claystone and coal seams. The lower

Narrabeen Group consists of a succession of sandstone units with interbedded sequences

of claystone and shale.

The niawarra Coal Measures crop out south of Clifton and generally form much of the

coastal plains in the Illawarra region, although they are frequently covered by

Quaternary alluvium.

The coal measures sequence is approximately 250 m thick extending downwards from

the base of the Coal Cliff Sandstone. The upper most layer is the Bulli Coal, which

consists of black bituminous coal of coking grade in a seam between 1.5 and 2 m thick.

Below the Bulli seam the sequence consists of carbonaceous shale and lithic sandstone

with interbeds coal sandstone and conglomerate lenses. Only the Bulli Coal has been

mined in this area.

The Narrabeen Group extends for the full length of the escarpment in the study area and

also crops out to the west of the escarpment. It is the principal group of rocks making

up the coastal cliffs between Scarborough and Stanwell Park where it is particularly well

exposed.

The lower Narrabeen Group contains the following units from top to base. The Bulgo

Sandstone consists of grey quartz-lithic sandstone with minor reddish-brown claystone


42

and some thin conglomerate bands. The Stanwell Park Claystone consists of red-brown

and greenish claystone containing two prominent quartz sandstone beds. The

Scarborough Sandstone consists of thickly bedded sandstone with some conglomerate

beds and rare shale. It outcrops boldly and forms the major part of the coastal cliffs.

The Wombarra Shale consists of grey shale, but it contains one thick sandstone layer

known as the Otford Sandstone Member. The Coal Cliff Sandstone is a massive grey

lithic sandstone which forms the coastal cliffs in the southern part of the study area.

3.2 ILLAWARRA COAL MEASURES


The niawarra Coal Measures overlie marine sediments of the Shoalhaven Group and are

in turn overlain by alluvial sediments of the Narrabeen Group in the Wollongong area.

Rocks of the Narrabeen Group crop out in the lower coastal cliffs and shore platforms,

extending as a narrow zone northeastwards from the base of the cliff north of Clifton.

South of Clifton the coastline is composed of Illawarra Coal Measures but the main

outcrop zone of the coal measures south of Wombarra is mainly located on the steep

slope at the base of the Illawarra escarpment. Here these strata are overlain by

extensive accumulations of talus, commonly in excess of 5 m thick.

The stratigraphy, depositional settings and petrography of the Late Permian Illawarra

Coal Measures have reported by Bamberry (1992). The Illawarra Coal Measures has

been divided into two subgroups. These are the Sydney and underlying Cumberland

Subgroups.

The Sydney Subgroup is some 90-130 m thick in the northern niawarra region and

consists of lithic sandstone, siltstone, claystone and coal with a minor amount of tuff.

Commonly along the outcrop of this unit the coal seams have been observed by the

author to be permeable and to act as aquifers. The Cumberland Subgroup is described

as being about 110 m thick beneath the northern Illawarra region where it consists of

sandstone with some interbedded claystone. The Shoalhaven Group underlies the
43

Illawarra Coal Measures and consists of sandstone and siltstone below the northern

Illawarra area, but includes latite members to the south which represent the Gerringong

Volcanic Facies.

3.2.1 UPPER ILLAWARRA COAL MEASURES (SYDNEY SUBGROUP)


The base of the Sydney Subgroup is marked by the base of the Wilton Formation

(Bamberry, 1992). The Sydney Subgroup in the Illawarra area contains the following

units.

3.2.2 WILTON FORMATION


The basal formation of the Sydney Subgroup is the Wilton Formation. It is 15 to 30 m

thick and comprises a basal coarse sandstone fining up to a predominantly siltstone

sequence. Typical examples of the sandstone sequence in the basal Wilton Formation

occur in the railway cutting at Thirroul (Fig. 1.3). It comprises very coarse-grained to

conglomeratic cross-bedded sandstone with a maximum thickness of about 1.5 m. The

Woonona Coal Member overlies the coarse basal sandstone and varies in thickness from

a few centimetres to greater 6 m (Bamberry, 1992) but it most commonly consists of

about 1.5 m of thinly bedded coal, carbonaceous shale and shale.

Higher sandstone beds within this formation are fine- to medium-grained and well

sorted. They exhibit a general increase in quartz content up through the formation and

northwards, and are classified as litharenite to sublitharenite. Matrix material consists

predominantly of authigenic illite, smectite, chlorite and kaolinite, with lesser

microcrystalline quartz. Commonly, the cement is calcite with lesser siderite. Siltstone

and shale in the Wilton Formation were examined using XRD analysis of whole rock

and clay fractions. The major components are quartz, illite, smectite and kaolinite

(Odins et al, 1990).


44

This formation is overlain by the Tongarra Coal (Bowman, 1974) which, in turn, is

overlain by the Bargo Claystone (Fig. 3.1).

3.2.3 TONGARRA COAL


This is the next persistent coal above the Woonona Coal Member. It consists of

interbanded dull and bright coal with interbeds of carbonaceous shale and tuffaceous

claystone. The Tongarra Coal has a regular thickness (Bamberry, 1992) of between 2.8

and 3 m in the coastal exposures. A tectonic control, namely the stable tectonic sett

of southern Sydney Basin, exerted a significant control on the distribution of the

Tongarra Coal (Bamberry, 1992).

3.2.4 BARGO CLAYSTONE


The Bargo Claystone typically consists of dark gray to black claystone, locally

containing minor sandstone (Austinmer Sandstone Member) and tuff (Huntley Tuffaceous

Claystone Member). This unit is commonly a soft, parallel laminated, pale claystone

which, in some cases, exhibits normal grading (Bamberry, 1992). The unit immediately

overlies the Tongarra Coal and is overlain by the Darkes Forest Sandstone.

The Bargo Claystone is about 15 m thick in the Illawarra region (Bowman, 1974), but

it ranges in thickness from less than 10 m to 38 m in the northeastern part of the

coalfield (Hutton et al, 1990). The Austinmer Sandstone Member is very fine- to fine-

grained sandstone and is moderately to well sorted. It is present near the base of th

formation interbedded with siltstone.

3.2.5 DARKES FOREST SANDSTONE

The Darkes Forest Sandstone is 10 m thick (Bowman, 1974), overlies the Bargo

Claystone and is immediately overlain by the Allans Creek Formation. The sequence

records the deposition of fine- to medium-grained sand, silt and clay in a delta fron
45

setting (Bamberry, 1992). This unit is predominantly a well sorted, fine-grained

litharenite. Matrix material is predominantly clay minerals and microcrystalline quartz.

The unit is cemented by coarse-grained calcite, with lesser siderite. Porosity within the

unit is very low due to this widespread cementation (Odins et al, 1990).

3.2.6 ALLANS CREEK FORMATION


The Allans Creek Formation directly overlies the Darkes Forest Sandstone and is

overlain by the Kembla Sandstone. Its thickness is usually between 4 and 16 m (Hutton

et al, 1990). This formation consists predominantly of fine-grained lithologies.

Typically, sandstone within the formation consists of moderately to well sorted, fine-

to medium-grained, subangular to subrounded chert grains, with lesser volcanic rock

fragments (Odins et al, 1990). Typical matrix materials include microcrystalline quartz

and clay minerals. The American Creek Coal Member in the upper part of the Allans

Creek Formation occurs immediately below the Kembla Sandstone.

3.2.7 KEMBLA SANDSTONE


This unit is situated between the American Creek Coal Member and the Wongawilli

Coal. The contact between the Kembla Sandstone and Allans Creek Formation is a

sharp erosional surface that is marked by abundant chert pebbles, intraclasts and

preserved fossil wood (Bamberry, 1992). The Kembla Sandstone is between 4 and 8 m

thick (Hutton et al, 1990). It is a well sorted, fine- to medium-grained litharenite.

Matrix material mainly consists of clay minerals. Porosity within this unit is low,

although where pores occur they are commonly very open in nature (Odins et al, 1990).

Two major lithofacies associations are recognised within the Kembla Sandstone. The

fining upwards sequence of the lower coarse member of the Kembla Sandstone, together

with the terrestrial aspect of surrounding strata attests to a fluvial origin for this

succession. The well ordered internal lamination within the overlying fine member
46

indicates a floodplain environment and suggests that the sequence was deposited in

meandering river setting (Bamberry, 1992).

3.2.8 WONGAWILLI COAL

Overlying the Kembla Sandstone is the 3 to 9 m thick Wongawilli Coal consisting of

coal, carbonaceous shale and tuffaceous claystone (Table 3.1). Coaly beds in the upper

and lower parts of the Wongawilli Coal are separated by tuffaceous claystone. The

upper coaly beds consists of carbonaceous claystone with coaly laminae whereas well-

developed coal, comprising interbanded bright and dull coal with bands of carbonaceous

mudstone and claystone, is found in the lower half of the unit.

The tuffaceous character of the claystone bed within the Wongawilli Coal indicates that

it represents an airfall ash deposit (Bamberry, 1992). In coastal outcrops, the basal

section of the Wongawilli Coal consists of a thick sequence of laminated carbonaceous

claystone and siltstone (2 to 4 m). At Scarborough (Fig. 1.3) the basal section overlies

overbank sequences of Kembla Sandstone with a low-angle erosional basal contact.

3.2.9 ECKERSLEY FORMATION


This formation overlies the Wongawilli Coal and is an interbedded sequence of fine- to

coarse-grained, quartz-lithic sandstone, siltstone, claystone and coal with tuff as an

important component of the clastic beds. Several units of member status, mostly coal

seams, are defined within the Eckersley Formation (Fig. 3.12). This formation generally

ranges 35 m thick to a maximum thickness of 112 m in the northern part of the

coalfield (Hutton et al, 1990). The Eckersley Formation is dominated by fine- to

medium-grained sandstone at the base. Overlying this sandstone is up to 14 m of

interbedded coal, carbonaceous siltstone and claystone of the Woronora Coal Member.

In cliffs near Scarborough and Clifton, Hanlon (1953) formally named the Hargrave,

Cape Horn and Balgownie Coal Members and the Lawrence Sandstone Member. The
47

respective outcrop type-section thickness of the Hargrave, Cape Horn and Balgownie

Coal Members are 0.46 m, 1.32 m and 1.30 m. The Lawrence Sandstone Member has

a maximum thickness of 18 m. The interval between the Balgownie Coal Member and

Bulli Coal normally comprises an upward fining coarse- to fine-grained fluvial sandstone

sequence capped by siltstone and shale (Loddon Sandstone Member; Bamberry, 1992).

This forms the uppermost clastic unit in the niawarra Coal Measures. It typically

occurs as a moderately to poorly sorted, medium- to coarse-grained litharenite with

minor feldspathic litharenite. Matrix material is commonly well-crystallised kaolinite

and less common illite and smectite. This sandstone has a carbonate cement and is

relatively non-porous.

3.2.10 BULLI COAL


The top of the Illawarra Coal Measures is clearly indicated by the Bulli Coal which is

2 to 3 m thick in the area of the detailed study. The unit can be seen to outcrop along

the waters edge from the Jetty Fault to Clifton Fault where it forms the weakest layer

in the lower cliff (Fig. 3.2). The original entrance to Coalcliff Colliery can still be see

on the rock platform just to the south of the Jetty Fault. North of the Jetty Fault, the

coal seam is believed to outcrop below sea level at the base of the Coal Cliff Sandstone

rock platform.

3.3 NARRABEEN GROUP


Narrabeen Group strata (Table 3.1) generally create significant problems for slope

stability within the niawarra region. When unweathered rocks are tested, the massive

sandstone and dense thinly bedded siltstone and shale members show high values for

most engineering parameters such as strength and durability (see chapter 7). However

weathered rocks show low values for most engineering parameters. Outcrop of
48

Narrabeen Group rocks tend to be deeply and intensely weathered, whilst even fresh

rock commonly deteriorates rapidly on exposure.

Constraints on the simple acceptance of engineering test data arise from the specific

lithologies in the Narrabeen Group. These include the common lithic nature of sand

grains in the sandstone units and the presence throughout the sequence of interbedded

claystone units which are prone to rapid weathering (Fig. 3.3). The presence of unstable

cements and swelling clays contribute to this rapid weathering and disintegration. In

addition, well-developed joints further reduce the overall rock mass quality.

A combination of these factors results in Narrabeen Group rocks commonly giving rise

to fall and slip-prone scree slopes and talus soils where the unit crops out on hillsides.

In sea cliffs, the Narrabeen Group strata are rather more easily eroded than the

Hawkesbury Sandstone strata, and hence relatively rapid rates of cliff-line recession

occur leaving the escarpment protected by the erosion resistant Hawkesbury Sandstone.

Differential weathering of the interbedded Narrabeen Group sequence, coupled with cliff-

line collapse and retreat has generated a benched topography on the Illawarra escarpment

(Fig. 1.4). As a consequence of their poor weathering characteristics, the rocks of the

group give rise to a disproportionate range of stability problems.

3.3.1 LOWER NARRABEEN GROUP


3.3.1.1 COALCLIFF SANDSTONE (CSs)
This is quite a competent unit being composed of fine- to medium-grained quartz lithic

and lithic moderately sorted sandstone. The unit forms a protective rock platform

extending from south of Coalcliff beach to the Clifton Fault at Moronga Park (Figs 3.7,

5.1). The unit is typically 10-12 m thick in the Coalcliff area near the Jetty Fault (Fig.

3.3). It consists of silt and clay laminae, normally exhibiting small-scale cross-

lamination parallel to irregular bedding (Fig. 3.3). Repeated 'fining upwards' sequences

are present within Coal Cliff Sandstone (Ward, 1972). Rockfalls are common in the
49

Coal Cliff Sandstone, especially between Clifton and Coalcliff, where undermining along

the shale bands reduces the support for the overlying vertically-jointed sandstone and

eventually leads to stabs falling off along the vertical joint faces (Fig. 3.3). The retreat

of the coastline is also brought about by rockfalls associated with the marine

undercutting of the Coal Cliff Sandstone (Fig. 3.2). The presence of the Bulli Coal

beneath this sandstone also enhances rockfalls since its permeability and aquifer

properties cause seepage concentrations and the rapid weathering of the underlying

siltstone, which breaks down to form clay typically possessing low shear strength

properties.

3.3.1.2 WOMBARRA SHALE (WSh)


This unit is troublesome in regard to slope stability along the entire escarpment. The

Wombarra Shale is typically 30 to 35 m thick in the area of the study and is composed

of greenish grey shale with fine-grained lithic sandstone interbeds and a thicker coarse

sandstone unit (Otford Sandstone Member). The Otford Sandstone Member has a

maximum thickness of 7 m with cross-bedding and a thin conglomerate bed (Fig. 3.5).

Undercutting occurs at the contact between Conglomerate sandstone and the cross-

bedded finer grained upper portion of the Otford Sandstone Member between Clifton

and Coalcliff along the Lawrence Hargrave Drive. The rapid weathering of the

interbedded siltstone along this contact produces a low shear strength clay while

indicates small rockfalls (Fig. 3.5). In the detailed study area, the formation is subject

to rapid weathering and erosion, with active marine erosion occurring along the coastal

cliff-line (Fig. 3.4). Weathering of the Wombarra Shale, which breaks down to form

low strength clay, and the existence of seepage concentrations within the fractured and

exposed surface layer cause instability along the coastal cliffs. Thus the Wombarra

Shale acts as a bedrock for many landslides along the Lawrence Hargrave Drive (Fig.

1.4). From the north headland at Coalcliff to the Clifton Fault the road is founded on
50

a small bench within this unit. This has led to numerous stability problems along the

road.

3.3.1.3 SCARBOROUGH SANDSTONE (SSs)


This unit forms the sheer cliff-line immediately adjacent to the coast extending from

north of Stanwell Park Beach to the north end of Coalcliff beach (Fig. 1.4). It also

crop out in the steep cuttings immediately above Lawrence Hargrave Drive from south

of Coalcliff to the Clifton Fault. The unit is typically 25 to 30 m thick and consists

predominantly of coarse-grained quartz lithic sandstone. Smaller scale 'fining upwards'

sequences are repeated within unit (Ward, 1972). This unit includes a few thin shale

beds and fairly common fine pebble to granular conglomerate beds. The presence of

interbedded claystone within the Scarborough Sandstone, which weathers to form low

strength clay, and the presence of vertical joints causes many rockfalls (Fig. 3.6). Sout

from the Coalcliff tunnel to the Clifton Fault it forms the foundation for the railway li

in a very unstable area (Rube Hargrave Park; Fig. 3.7). Below the railway line and

above Stony Creek the outcrop of the Scarborough Sandstone is completely obscured

by steeply sloping talus detritus down to the level of the Wombarra Shale (cf. Fig. 1.5).

It would underlie the road at Stony Creek crossing if it had not been eroded away.

3.3.1.4 STANWELL PARK CLAYSTONE (SPC)


The Stanwell Park Claystone is approximately 36 to 40 m thick in the area of the

detailed study. It is made up of interbedded green to purple/brown claystone and minor

quartz lithic sandstone. Hanlon (1953), in his section of the Stanwell Park Claystone,

recorded numerous sandstone beds up to 4 m thick and estimated the claystone to

sandstone ratio as about 2 to 1. The land slides referred to by Shellshear (1890)

between Stanwell Creek and Coalcliff Railway Station were associated with this

formation. The author has observed several landslides within the Stanwell Park
51

Claystone in the State Rail Authority area. It forms the foundation for the railway line

between Coalcliff Station and the Coalcliff tunnel (Fig. 3.8). The rapid weathering of

the StanweU Park Claystone unit, which breaks down to form clay typically possessing

low shear strength properties and the existence of seepage concentrations within the

fractured surface layers of the Stanwell Park Claystone cause instability in the study

area. Many of the slides associated with the Stanwell Park Claystone have other

complicating causes and are discussed later.

Fractured zones within this unit also appear to act as outlet points for subsurface water

which is believed to percolate down through joints and fractures in the overlying strata.

This unit is quite prone to rapid weathering and is responsible for the creation of the

relatively flat benches within the slopes of the escarpment.

3.3.1.5 BULGO SANDSTONE (BSs)

The Bulgo Sandstone is the thickest of formation in the Narrabeen Group in the

Illawarra area. It is 120 m thick at Clifton, forming over half the total Narrabeen

succession (Ward, 1980). The unit crops out at the top of the escarpment cliffs between

Clifton and Coalcliff and at sea level in the coastal headlands to the north of Stanwell

Park.

The Bulgo Sandstone in coastal districts can be divided into three distinct facies (Ward,

1980). The lower part is very similar to the underlying Scarborough Sandstone and it

was called "pebbly facies" by Ward (1980). Overlying the pebbly facies is a succession

of sandstone, shale, and granule conglomerate which has a very characteristic green

colour in the field. Because this green sediment consists of altered intermediate to basic

volcanic material it is referred to as the "volcanic facies". The sequence between the

top of the volcanic facies and the base of the Bald Hill Claystone has a considerably

higher proportion of shale than the remainder of the Bulgo Sandstone. This part of unit

was called "shaly facies" by Ward (1972).


52

The sediments of the Bulgo Sandstone are fluvial deposits (Ward, 1972). The varying

grain size and the proportion of shale reflect the relative intensity of river action

involved.

This unit is typically quartz lithic sandstone with mainly a medium to coarse grain size,

even conglomeratic in part. Smaller scale 'fining upwards' sequences are repeated

within Bulgo Sandstone (Ward, 1980). Interbedded bands of claystone are present at

the bottom and top of the unit. The presence of interbedded claystone within the Bulgo

Sandstone, which weathers to form low strength clay and undercuts the contact between

sandstone and claystone, causes many small rockfalls along the Lawrence Hargrave

Drive between Coalcliff and Clifton (Fig. 3.9).

In the study area, the Bulgo Sandstone was found to be approximately 120-125 m thick.

It is well jointed and forms the steep slopes below the Hawkesbury Sandstone cliff-

line. High plasticity clay bands exist within the unit which have been found to

contributed to slope stability problems in the area.

3.4 IGNEOUS ROCKS


Igneous activity is limited to small intrusions in the sedimentary rocks. Mining

operations in the area have provided valuable information on dykes and sills. The rocks

forming the dykes and sills are basalt or lamprophyre. Colliery investigation boreholes

in the Stanwell Park valley, to the east of the Illawarra Railway, are reported to have

detected intrusive sills in the Bulli Coal. It is reported that this has resulted in the coal

deposit becoming cindered and thus undesirable for extraction. These data are supported

by the exposure of a fresh basaltic sill which intrudes the Scarborough Sandstone about

700 m north of Stanwell Park Beach (Fig. 1.3).

The plans of the mine workings clearly display the multitude of dykes in the area, the

directions of which are highly dependant upon the jointing and faulting. A s a result,
53

the majority of the intrusions run in a north-south direction, with the remainder typically

following the minor maxima of 055 and 155.

F e w dykes are visible on the ground surface in the study area. O n e is visible in the

Scarborough Sandstone cliff-line some 30 m south of Stanwell Park Beach. It is

believed that this dyke extends in a southwest direction influencing the shape of the

upper headland. The second exposure, which is actually a dyke zone can be seen some

150 m upstream from the Stanwell Creek railway viaduct. This zone is about 15 m

wide and contains three separate dykes up to 1 metre wide (Adamson, 1974). A third

exposure is a weathered dyke mentioned by (Hanlon, 1958) on the road between

Coalcliff and Clifton. The presence of dykes is a contributing factor for slope

instability, but their effect needs to be determine with more exploratory work. The

dykes can affect the underground water flow and m a y increase seepage. The properties

seem to vary depending upon the contact strata. Observations by Harper (1915) and

A d a m s o n (1974) suggested that strong alteration of the intrusions occurs when they are

in contact with coal. This is confirmed by many of the dykes reported in the mine

workings consisting of green clay. If not in contact with coal, the dyke material m a y

exist as a fresh, hard, unaltered basalt or lamprophyre.

3.5 TALUS

M u c h of the bedrock exposures are blanketed with a talus mantle of variable thickness

and nature. Talus thickness along the escarpment ranges from zero to 20 m . Drilling

and excavations conducted by the State Rail Authority in the course of the major slip

remedial works from 1988-1992 revealed typical thicknesses of 5-10 m in the detail

study area.

Multiple massive talus deposits were observed which appear to have resulted from

ancient landslides. The properties of the talus are highly dependent upon its source.

Coarse talus is derived from sandstone strata and commonly consists of very large
54

boulders within a sand clay matrix (Fig. 3.10). Talus derived from claystone is often

finer grained having a larger clay fraction and higher plasticity index. This often occurs

at the base of the talus deposits which commonly mantle the coastal terraces developed

on the shale units. As a result, in a landslide situation, residual shear strengths often

reflect the properties of the clays which are derived from the basal claystone strata.

3.6 SEDIMENTARY STRUCTURES


Sedimentary structures are very common in the Illawarra Coal Measures and Narrabeen

sequences. Brief mention is made of them here because of their influence on the

escarpment stability. Some units within the niawarra Coal Measures are characterised

by the presence of flat bedding and burrowing.

The Narrabeen Group sandstone is characterised by ripple marks and planar high-angle

and trough cross-beds. Shale is characterised by laminated. The claystone units trend

to be structureless other than for some graded bedding. The Wombarra Shale and

Stanwell Park Claystone usually lacks structures apart from bedding. The larger

sedimentary structures result in anisotropy in many of the rocks. However, the

structures are always cemented, thereby reducing their effect on the variation in

orientation of mechanical properties.

3.6.1 SEDIMENTARY ENVIRONMENTS


The depositional settings of the Late Permian Illawarra Coal Measures are deltaic to

fluvial system (Bamberry, 1992). The sedimentary environment of the Narrabeen Group

in the southern part of the Sydney Basin indicates several different type of fluvial

deposits, with an upward succession from piedmont conditions with braided streams to

a swampy deltaic plain (Ward, 1972). This sequence is interpreted as the onshore

portion of a slow marine transgression, probably brought about by subsidence coupled

with declining erosive activity in the hinterland.


55

3.7 SUBSURFACE GEOLOGICAL SEQUENCES AND STRUCTURES

RECOGNISED IN DRILL HOLES

Kembla Coal and Coke (KCC) have carried out a drilling program associa

planned development in the West Cliff and North Cliff Collieries. Bore

drilled to some 500 m depth, with the lower 30 to 40 m being fully cor

(1) 20 m of roof rock above the coal;

(2) the BuUi seam; and

(3) the immediate floor to about 10 m below the seam.

Some of the boreholes have been fully cored through the Narrabeen Grou

This part of the chapter presents the lithology of the upper coal meas

Narrabeen Group in the fully cored boreholes IL55, IL57, IL60 and IL64

West Cliff area at the locations shown in Figure 3.11.

3.8 GEOLOGY AND SEDIMENTARY STRUCTURES


3.8.1 UPPER ILLAWARRA COAL MEASURES
The upper niawarra Coal Measures, which range in thickness from 10 m t

consists of Eckersley Formation (Hargrave seam, Cape Horn seam, and La

Sandstone, Balgownie Coal and Loddon Sandstone Members) and Bulli Coal

IL57 (Fig. 3.12b). Only the upper part of this sequence above the Balg

intersected in boreholes IL60 (Fig. 3.12a) and IL64 (Fig. 3.12c). The

Eckersley Formation is 27.8 m thick and comprises coal (Hargrave and Ca

claystone, coaly claystone, carbonaceous claystone and fine-grained sa

IL57 (Fig. 3.12b). Lawrence Sandstone Member is 8 m thick and contains

medium-grained sandstone with ripple bedding at borehole IL57 (Fig. 3.1

Balgownie Coal is about 0.8 m thick in each borehole. The thickness of

medium-grained Loddon Sandstone Member is more variable being 5.7 m th


56

borehole IL57, 7 m at borehole IL60 (Fig. 3.12a) and 6.4 m at borehole IL64

(Fig. 3.12c). The Bulli Coal at the top of the niawarra Coal Measures is 2.8 m thick

at borehole IL60, 2.3 m thick at borehole IL57 and 1.90 m thick at borehole IL64.

3.8.2 COAL CLIFF SANDSTONE


The Coal Cliff Sandstone, which ranges in thickness from 6 m to 26.6 m, consists of

interbedded sandstone and siltstone with conglomerate occurring locally. Sandstone is

dominant in the four boreholes (Fig. 3.13). In borehole IL57 (Fig. 3.13b) this unit

consists largely of fine- to medium-grained sandstone, with conglomerate at the base of

the formation. In boreholes IL60 (Fig. 3.13a), IL64 (Fig. 3.13c) and IL55 (Fig. 3.13d)

sandstone is dominant throughout ranges from fine- to coarse-grained and is massive or

cross-bedded. In some beds is the cross-bedding poorly defined. The sandstone

commonly shows an upward decrease in grain size. It is overlain by shale or claystone.

The basal part of this unit consists of dark grey to black claystone or carbonaceous

claystone (boreholes IL55, IL60 and IL64). The dark colour results from the preserved

organic matter in these claystone units. Lower part of the unit contains several fractures

at intermediate angles to the core which have well developed slickensides. The

gradational arrangement of the coarse and fine members suggest that the sandstones

were deposited as channel-fill successions.

3.8.3 WOMBARRA SHALE


The Wombarra Shale ranges in thickness from 26 to 49.5 m. This unit consists of

mostly shale and claystone with interbedded sandstone and conglomerate occurring

locally (Otford Sandstone Member; Fig. 3.14). The ratio of sandstone to shale/claystone

varies from one borehole to another. In boreholes IL55 (Fig. 3.16d), IL57 (Fig. 3.16b)

shale/claystone dominates whereas in borehole IL64 (Fig. 3.16c) this unit consists of

subequal quantities of sandstone and claystone.


57

The sandstone in this unit is dominantly fine- to medium-grained and it commonly

shows an upward decrease in grain size. The basal part of this unit consists of dark

grey shales and interbedded very fine- to fine-grained sandstone. The dark colour

resulted from the preserved organic matter in some shale/claystone beds. The interbedded

very fine- to fine-grained sandstone shows ripple bedding IL60 (Fig. 3.16a). The green

to grey colour through most of the shale/claystone was related to the paucity of

preserved organic matter in it. Fissility in Wombarra Shale is very clear (Fig. 3.15).

The coarse members represent channel-fill or crevasse splay deposits and the fine

members are overbank and floodplain sediments.

3.8.4 SCARBOROUGH SANDSTONE


The Scarborough Sandstone ranges in thickness from 37-50 m and consists of mostly

sandstone and conglomerate with interbedded claystone and shale. Sandstone is the

dominant lithology in the four boreholes. In boreholes IL60 (Fig. 3.17a), IL57 (Fig.

3.17b) and IL64 (Fig. 3.17c) this formation consists largely of sandstone, especially in

borehole IL64. The sandstone in this formation is dominantly fine- to very coarse-

grained. Most of the sandstone is massive but some shows cross-bedding, a prominent

vertical joint with a length about 1 m is present in the sandstone at borehole IL55 (Fig.

3.18). The grey claystone/shale interbeds are usually thin (about 0.5-0.7 m ) . The grey

colour of the claystone/shale is related to the paucity of preserved organic matter in

them. The basal part of formation consists of conglomerate (boreholes IL55, IL57) or

fine- to medium-grained sandstone (boreholes IL60, 1164) which overlies the Wombarra

Shale. The fining upwards sequences in the Scarborough Sandstone suggest a fluvial

origin for this succession.


58
3.8.5 STANWELL PARK CLAYSTONE
The Stanwell Park Claystone ranges in thickness from 12-19.5 m. It consists mainly

of claystone and siltstone, with sandstone and conglomerate occurring locally (Fig.

3.19a). The ratio of claystone to sandstone varies from one borehole to another. In

boreholes IL55 (Fig. 3.19d), IL57 (Fig. 3.19b) and IL64 (Fig. 3.19c) claystone is

dominant whereas in borehole H64 this formation consists largely of sandstone with

minor claystone mainly at the base. The sandstone in this formation ranges from fine-

to coarse-grained and commonly shows an upward decrease in grain size. The basal

part of this unit consists of dark greenish-grey and reddish-brown to purplish claystone.

The dark colour results from the preserved organic matter in greenish-grey claystone

beds. The coarse members are probable channel (IL60) or crevasse splay deposits

whereas the fine members represent overbank and floodplain sediments.

3.8.6 BULGO SANDSTONE


The Bulgo Sandstone ranges in thickness from 154-167 m. It is extremely sandy in the

boreholes studied. In boreholes IL57 and IL64 sandstone accounts for about 85% of

the unit. The majority of the sandstone shows cross-bedding and ripple bedding.

Sandstone beds in the upper of the formation are generally finer than those in the lower

part of the formation and claystone is slightly more abundant. The basal part of

formation consists of fine- to coarse-grained sandstone in the four boreholes (Fig. 3.20).

At the base of formation dark greyish-green coarse- to very coarse-grained sandstone

containing abundant lithic grains has a sharp erosional contact with the underlying

Stanwell Park Claystone. This lower part is very similar to the underlying Scarborough

Sandstone and represented the "pebbly facies", of Ward (1980).

The succeeding 51 m of this unit consists of fine- to very coarse-grained sandstone

interbedded with dark grey shale, grey claystone and very coarse granule conglomerate.

The thickness of interbeds ranges between 0.5-0.8 m for shale, between 0.02-1.7 m for
59

claystone, between 0.04-1.0 m for fine siltstone and between 0.07-0.75 m for

conglomerate (Fig. 3.20). It is referred to as the "volcanic facies" (Ward, 1980).

The fine- to medium-grained sandstone contains prominent vertical joints with a length

of between 0.7-1.5 m and subvertical joints with a maximum length of 1.5 m (e.g. Fig.

3.21). The upper part of the unit contains more greyish-grey claystone (Fig. 3.20a) and

represents the "shaly facies") of Ward (1980). It was overlain by the Bald Hill

Claystone. The fining upwards sequences in the Bulgo Sandstone suggest a fluvial

origin for this succession.

3.9 DISCUSSION

The importance of stratigraphy in slope development was emphasised previously.

Stratigraphy must be determined in considerable detail for stability evaluations or design

investigations. All stratigraphic work and sample logging should be oriented toward

engineering applications. Particular attention should be directed toward recognition and

description of (a) thick sequences of weak rocks (claystone and shale) and the nature

of interbedded sequences, (b) thin marker beds, and (c) old failure surfaces.

In the Narrabeen Group, thick sequences of weak rocks (Stanwell Park Claystone and

Wombarra Shale) are rather more easily eroded than sandstone strata and hence

relatively rapid rates of recession occur. Undermining along these thick sequences of

weak rocks, at the contacts between the claystone and sandstone, reduces the support

for the overlying vertically-jointed sandstone and eventually leads to stabs falling off

along the vertical joint faces. Thin marker beds (coal seams) in the niawarra Coal

Measures commonly act as aquifers, with claystone beds acting as aquitards. Slope

instability is usually related to the presence of the aquifers which are the source of high

pore water pressures.


60

The lithic nature of sand grains in the sandstone units, the presence throughout the

sequence of interbedded claystone units which are prone to rapid weathering, and the

weU developed joints all tend to reduce the overall rock mass quality.

Landslides that occur within in soft rock masses (Stanwell Park Claystone and

Wombarra Shale) typically involve bedding as the basal shear surface, and joints or

other defect controls for lateral and headscarp release (Brown, 1974; Hancox, 1974).

According to Bell and Pettinga (1988) two types of bedding control can be identified:

(a) the presence of clay-rich interbeds, such as claystone within the Bulgo Sandstone,

and (b) the presence of organic-rich sediments, such as carbonaceous claystone in the

Illawarra Coal Measures. The geotechnical significant of bedding control is two-fold:

firstly by providing a layer of lower permeability above which positive pore water

pressures can increase; and secondly the presence of relatively low shear strength

materials within which or upon which sliding may occur (Bell and Pettinga, 1988).

According to Barton (1988) sedimentological control has an affect on slope stability by

(a) changes in lithology causing a permeability contrast allowing localised high pore

pressures; (b) increased clay content - this was suggested by Simpson and Walton

(1970) as the initial cause of most continuous bands of clay mylonite observed in the

English coal measures; (c) changes in clay mineralogy - this has been suggested by

Fookes (1967) for shear surfaces in the Siwalik Clay; and (d) dissolution of soluble

minerals resulting in increased porosity and lower shearing resistance.

In the Coal Cliff Sandstone along the coast-line between Clifton and Coalcliff the fine-

grained sandstone near the base of the formation sometimes grades into claystone (Fig.

3.22a). Weathering and erosion of this claystone reduces the slope stability on the Coal

Cliff Sandstone. In some places, near the top and middle of this formation, claystone

interbeds occur within the dominantly sandstone units, and result in the stability of the

sandstone unit being decreased by the water flow and weathering of the claystone

(Fig. 3.22b).
61

Fractured zones within Narrabeen Group rocks appear to act as outlet points for water

percolation. Commonly the Scarborough Sandstone is conglomeratic with chert clasts,

and shows moderately porosity and permeability especially in the basal half. This has

produced a perched water table in this formation, with water flow concentration along

the junction with the underlying Wombarra Shale. Weathering and erosion of the

Wombarra Shale causes many rockfalls (Fig. 3.23). Between Clifton and Coalcliff along

the Lawrence Hargrave Drive some sandstone beds are present within the Wombarra

Shale. The sandstone beds are more prominent and have the effect of reducing the

amount of weathering of the shale that is exposed and eroded away. It has been

observed at this locality, and at other localities throughout the study area, that often

when interbeds of a more competent rock type occur within a claystone unit, they have

the effect of stabilising the claystone unit. In contrast, whenever claystone interbeds

occur within dominantly sandstone units, the stability of the sandstone unit is decreased.

In considering the Narrabeen Group rocks, the Wombarra Shale and the Stanwell Park

Claystone are generally least influenced by the stability afforded by sandstone interbeds.

The greatest stability for sandstone unit occurs when there is a low thickness of

claystone interbeds. Amongst the Narrabeen Group sandstones, the Bulgo Sandstone is

likely to be the most unstable because this formation consist of thickly bedded sandstone

with numerous intercalated siltstone and claystone beds up to 3 m thick. This

interbedded claystone has most effect on the stability of formation. Weathering and

erosion of claystone causes many rockfalls in the study area especially between Clifton

and Coalcliff along the Lawrence Hargrave Drive (Fig. 3.9b). In contrast the most

stable unit is the Scarborough Sandstone which overlying a few lenses of claystone.
62
63

CHAPTER 4

PETROLOGY OF NARRABEEN GROUP SANDSTONE

4.1 INTRODUCTION
McElroy (1954) studied the composition and texture of the Narrabeen Group sediments

in a systematic manner. Loughnan (1963) presented a petrological study of a vertical

section through the Narrabeen Group at Helensburgh. Ward (1971a,b) made a

comprehensive petrological study of the Narrabeen Group sediments in the southern

Sydney Basin. Bai (1991) studied the petrology, diagenesis and reservoir potential of

the Narrabeen Group sandstone in the Sydney Basin. All these studies show that the

content of detrital quartz grains generally increases towards the top of the group and

that the detrital lithic fragment content generally decreases.

The aims of this petrological study are to determine variations in the detrital

composition of the sandstone units, and to assess the effects of these compositions on

weathering, engineering properties and slope stability in the northern niawarra. The

influence of composition of sandstone on engineering properties of the Narrabeen Group

and land instability will be discussed in chapters 7 and 8.

4.2 STUDY METHODS


Thin sections were cut from 40 sandstone samples collected from the geologically

logged borehole DL55 in the North Cliff area and outcrops between Scarborough and

Stanwell Park Station. They were impregnated with blue dye to facilitate observation

of porosity in thin sections. All thin sections were examined under a petrological

microscope to observe grain size, mineralogy, texture, porosity, dissolution features,

sorting and roundness. Some samples are altered to such an extent that the detrital

lithic fragments could not be distinguished from the clay matrix and such samples are
64

considered to be unsuitable for reliable point counting. Of the 40 samples, a total of

36 thin sections were selected for point counting. Photomicrographs were taken of

selected thin sections using an Olympus (Model BHSP) Camera with 100 ASA film.

Except in a few cases, at least 250 counts were made for each slide to ensure a

reasonably high accuracy in the result. The results of point counting are presented in

Tables 4.1 to 4.4.

Sorting and rounding were estimated from thin section examinations. The degree of

sorting ranges from very poorly, through poorly to well sorted and the degree of

rounding ranges from angular, through subangular and subrounded to well rounded.

Even though porosity characteristics were examined under the petrological microscope,

the study of porosity characteristics is not an easy task since the blue dye penetration

is not always complete in every thin section.

4.3 MINERAL COMPOSITION


4.3.1 QUARTZ
The detrital quartz grains recognised consist of three types:

(1) unstrained monocrystalline grains displaying uniform extinction;

(2) strained monocrystalline grains displaying undulatory extinction; and

(3) polycrystalline grains.

In addition, a trace amount of chalcedony was also recognised in a few thin sections.

Usually fresh samples have the highest quartz content, whereas weathered samples have

the lowest quartz content.

It seems that at least some of the quartz grains were derived from granite (Blatt et al,

1980). Some of the detrital monocrystalline quartz grains contain many fluid inclusions

and have a very cloudy appearance under plane polarised light. They are most likely

derived from hydrothermal vein sources (Fig. 4.1).

Detrital polycrystalline quartz grains are present in three different forms:


65

(1) composite grains consisting of uniformly sized crystals, which have mostly straight

or sutured contacts;

(2) composite grains consisting of different sized crystals; and

(3) composite grains with preferred orientation of elongate crystals, some of which

have sutured contacts. Type (3) is the least abundant.

The detrital polycrystalline quartz grains (Fig. 4.1) are considered to be largely derived

from metamorphic rocks, including quartzite, as most of them contain more than 5

individual crystals, sutured contacts and or preferred orientation (Blatt et al, 1980). In

addition, a small number of the polycrystalline quartz grains show a cloudy appearance

under plane polarised light, which is caused by abundant fluid inclusions. Quartz

overgrowths were recognised in a number of thin sections (e.g. Fig. 4.5), particularly

those rich in quartz, and in some thin sections it is not easy or possible to distinguish

the quartz overgrowth from the detrital quartz grain due to lack of a clear boundary.

For this case cathodoluminescence is used to determinate the quartz overgrowth from

the detrital quartz grain.

4.3.2 FELDSPAR

The majority of the sandstones, especially those rich in detrital quartz grains, have less

than 2% detrital feldspar. Detrital feldspar grains are generally altered. In fact, some

of them have altered to such an extent that it is impossible to identify the feldspar. In

addition detrital feldspar grains are easy targets for carbonate replacement (Fig. 4.2).

Some have been partly or completely replaced by carbonate and some of the secondary

pore spaces in thin sections were probably created by the dissolution of detrital feldspar

grains.

Patchy kaolinite flakes, filling the outline of detrital feldspar grains; are considered to

result from replacement.


66
4.3.3 ROCK FRAGMENTS
Rock fragments include volcanic, metamorphic and sedimentary rock fragments.

Volcanic rock fragments mostly comprise chert grains that may have been derived from

silicic volcanic rocks or from silicified flows and tuff. Sedimentary clasts consist

mudstone, shale and siltstone fragments. Metamorphic rock fragments consist mainly

of slate and schist and were recognised by the preferred orientation of elongate min

in the fragment. Rock fragments make up more than 50% in some of the lithic

sandstones (cf. Boggs, 1992).

4.3.4 CHERT

Chert is made up of microcrystalline quartz (Midgley, 1951) as determined by X-ray

diffraction. It is distinguished from detrital polycrystalline quartz grains by its

size, and from volcanic rock fragments by its relatively clean surface under the

petrological microscope and its quartz composition. The distinction between silicifi

volcanic chert and detrital chert grains is not possible in some thin sections. In t

study, the uncertain detrital clasts were counted as chert. In some chert grains,

radiolarian remains, which occur as nearly perfect spheres filled with quartz, were

recognised. Their presence provides strong evidence for the identification of the gr

as detrital chert.

4.3.5 MICA

Mica includes muscovite and biotite with muscovite being more abundant. Due to

mechanical compaction, elongated muscovite plates are usually deformed around the

adjacent hard detrital grains. Mica content ranges from 0.2% to 3.6% of the total

sandstone.
67
4.3.6 ACCESSORY MINERALS

In the majority of samples, no accessory minerals were identified. Where present they

include tourmaline, zircon and opaque grains. In the samples with accessory minerals,

the content is generally 0.2% to 1%.

4.3.7 IRON OXIDES

Iron oxides are present in the all of the samples. They range from 0.6% to 14.5% of

the total sandstone. They are determined by the brownish colour, irregular shape,

opaqueness and aphanitic texture. They occur both as detrital grains and as secondary

cement.

4.3.8 CARBONATE

Carbonate occurs in two types of texture: microcrystalline and macrocrystalline, with the

latter as the dominate one. The former usually occurs as irregular patches filling inter-

grain pore spaces. The macrocrystalline carbonate crystals generally show two clear

cleavage directions and have clean surfaces under plane polarised light. The Carbonate

not only fills inter-granular pore spaces but also partly or completely replaces/corrodes

detrital grains, as well as forming a coat on detrital grains. Carbonate cement coating

detrital grains is much less common than cement filling inter-granular pore spaces and

the secondary carbonate replacing/corroding detrital grains.

The majority of the sandstone samples contain carbonate cement. The highest content

of carbonate (carbonate cements and replacement carbonate) was recorded as 35.6% of

the total sandstone in a fresh core sample from the Coal Cliff Sandstone (Fig. 4.2). In

some thin sections, partially dissolved carbonate cements were recognised. The

dissolution has created secondary pore spaces. In the point counting no attempt was

made to record the components replaced by carbonate (mainly feldspars and volcanic

rock fragments). Therefore, the content of carbonate is high.


68

The carbonate cements include calcite, siderite and perhaps ferroan calcite, dolomite and

ankerite. However, the differentiation of them is not easy under the petrological

microscope.

4.3.9 KAOLINITE

Kaolinite occurs in the all of the samples, ranging from 0.6% to 31.3% of the total

sandstone. Kaolinite flakes filling inter-granular pore spaces can often be identified

positively under the petrologic microscope (Fig. 4.3) and by scanning electron

microscopy (SEM; Fig. 4.4a). It is generally the product of intense weathering of

feldspar or clayey rocks in warm and humid climates; kaolinite is often difficult to

distinguish from chert and finely crystalline volcanic rock fragments. X-ray diffraction

provides the most certain method of identification.

4.3.10 MATRIX

The definition of matrix used here follows that of Morris et al. (1979). It generally

refers to all particles which are too fine to be identified with the petrological

microscope. Thus it includes all clay minerals, which may be detrital or diagenetic, clay

(< 4 microns) and silt-sized (4-62.5 microns) mica, quartz and feldspar, and the products

created by breakdown of fine-grained rock fragments.

The definition of clay fraction in soil mechanics is different. In the latter definition c

fraction is that percentage of soil which has a particle size below 0.002 mm (2 microns)

and which consists of clay minerals. The fraction below 0.002 mm which does not

consists of clay minerals is called "clay-size" fraction and not clay fraction. Silt sized

particles are not included in either definition.

The matrix commonly occurs as patches filling inter-granular pore spaces. The most

common matrix observed in the thin sections consists of detrital clay minerals and silt-

sized quartz grains. Most of the sandstones studied have a matrix content forming more
69

than 10% of the total sandstone. The highest matrix content of 28.6% was recorded in

a slightly weathered Bulgo Sandstone (Table 4.3b, sample SWBSs6). The point count

data show that the amount of matrix is generally inversely related to the grain size.

The coarse-grained sandstones tend to have less matrix than the fine-grained sandstones.

In addition, the amount of matrix generally increases at the same time as the volcanic

rock fragment context decreases.

The quantities of matrix and carbonate cement are also inversely related. The higher

the carbonate cement, the lower the clay matrix. This is illustrated in Table 4.4b where

the Otford Sandstone Member conglomerate sample (SWCONG5) has a much higher

carbonate content than the matrix-rich sandstone sample (SWOSM4).

4.3.11 CEMENT

The majority of the sandstone samples contain calcite, kaolinite and iron oxide cement,

but minor chlorite and quartz cement is present in some samples (e.g. Fig. 4.4b).

Calcite and kaolinite are the most common cementing agents in the Coal Cliff

Sandstone. Calcite, siderite and kaolinite cements occur in the sandstones of Wombarra

Shale (Otford Sandstone Member). Kaolinite and calcite are common cementing agents

for the Scarborough and Bulgo Sandstones (Figs 4.5, 4.6), but minor quartz cement is

also present in some samples.

In some thin sections, authigenic chlorite coating detrital grains can be identified

positively under the petrologic microscope.

4.3.12 POROSITY

Since all thin sections were impregnated with blue dye, the pore space were identified

by the blue colour under plane polarised light. Porosity includes pores of both primary

and secondary origin. Visual porosity varies from 0.2% to 4% of the total sandstone.
70

4.3.13 TEXTURE OF SANDSTONES


As indicated by point count data, the Narrabeen Group sandstone in the northern

Illawarra between Scarborough and Stanwell Park consists principally of detrital quartz

grains, lithic fragments (including chert) and detrital feldspar grains.

The rock fragments range in roundness from angular to well rounded with the majority

of them being subangular to subrounded. The well rounded quartz grains probably

represent second cycle sediment and were derived from earlier deposited sandstone

sequences. The poorly rounded angular quartz clasts were derived from volcanic

sources.

The sandstone of the Narrabeen Group in this area ranges in sorting from very poorly

to well sorted. Coarser sandstones tend to be poorly to very poorly sorted whereas finer

sandstones tend to be moderately to well sorted.

4.3.14 CLASSIFICATION OF SANDSTONE

The classification of sandstone is a controversial subject. A summary of sandstone

classifications was given by Pettijohn et al (1987). The sandstone classification (Fig.

4.7) used in this thesis is modified from the one proposed by Folk (1980). A 50%

quartz line is added to Folk's classification to emphasise those sandstones with more

than 50% but less than 75% detrital quartz grains. The adjectival modifier "quartzose"

is used to differentiate these lithologies, e.g. quartzose litharenite. The results of

calculations for the classification of Narrabeen Group sandstones are shown in Tables

4.5 to 4.8. Based on this classification fresh Coal Cliff Sandstone is quartzose

litharenite whereas slightly weathered to moderately weathered sandstone generally falls

in the litharenite field (Fig. 4.7a). All of the samples are moderately sorted. Fresh and

slightly weathered samples from the Otford Sandstone Member are litharenite and

quartzose litharenite (Fig. 4.7b). Fresh, slightly and moderately weathered Scarborough

Sandstone samples are all quartzose litharenite (Fig. 4.7c). They are moderately sorted
71

sandstones. Fresh Bulgo Sandstone ranges from litharenite to sublitharenite. Slightly

weathered and moderately weathered samples are quartzose litharenite (Fig. 4.7d). These

samples are moderately sorted.

4.4 PETROLOGICAL AND MINERALOGICAL ASPECTS OF WEATHERING


Thin sections have been cut from fresh and weathered sandstone from the Narrabeen

Group so that the effects of alteration with an increase in weathering could observed.

Forty thin sections were examined, however, only those thin section which show specific

relevant features are described. The rocks have been sampled from a variety of

locations. Fresh samples came from borehole IL55 at different depths. This borehole

was drilled by Kembla Coal & Coke Pty Ltd ( K C C ) in the North Cliff Collieries.

Table 4.1 shows the analyses of fresh, slightly weathered and moderately weathered

samples from the Coal Cliff Sandstone. Weathered samples were selected between

Clifton and Coalcliff and fresh samples came from borehole IL55 at depths of 435.75 m

to 442.3 m .

Table 4.4 shows analyses of fresh and slightly weathered samples from the Otford

Sandstone Member. Fresh samples came from borehole IL55 at depths of 411.5 m to

414.7 m. Slightly weathered samples were selected from the side of the road just south

of Jetty Fault between Clifton and Coalcliff.

Table 4.2 shows analyses of fresh, slightly weathered and moderately weathered samples

from the Scarborough Sandstone. Weathered samples were collected between Clifton

and Coalcliff. Fresh samples came from the borehole IL55 at depths of 394.7 m to

402.1 m.

Table 4.3 shows analyses of fresh, slightly weathered and moderately weathered

samples from the Bulgo Sandstone. Fresh samples came from the IL55 borehole at

depth of 289.6 m to 316 m. Slightly weathered samples were selected between Clifton,

and Coalcliff, while moderately weathered samples came from Stanwell Park Station.
72

It was not necessary to make thin sections of all the highly weathered specimens; X-

ray diffraction was a valuable tool in determining the mineralogy of these specimens.

These specimens have been weathered at the surface. Specimens have a distinct milky

appearance indicating that the fresh constituents of the rock show some degree of

alteration, especially the feldspar and rock fragments. The progressive increase in iron

oxide content and related decrease in siderite content is a result of the soluble ferrou

ion being unstable in an oxidising environment. With a fluctuating water table the

ferrous ion from the siderite is converted into the insoluble ferric ion to form iron

oxides, usually limonite. It appears that the formation of the iron oxides is accompanied

by a wedging action which prises some grains apart and thus reduces the perfect

packing. Decrease in chlorite content is evident with an increase in weathering (Figs

4.8 to 4.10).

Some samples have been weathered in a non-oxidising environment because they were

in a location where alteration was due to the effect of groundwater flow. This

groundwater was probably in a confined aquifer which was in a reducing environment.

The coarse-grained nature of the sandstone would have undoubtedly aided the passage

of water and thus weathering. In strongly ferruginised moderately weathered sandstone

most of the matrix and cement has been replaced by iron oxides, and rock fragments

have been partially altered.

From all the thin sections examined, some general conclusions have been drawn

concerning the changes occurring due to an increase in weathering from fresh to

weathered strata. These are listed below :

(1) Decrease in siderite and calcite content with an increase in weathering (Fig. 4.11).

This is caused by dissolution of the carbonate cement once it comes into contact

with bicarbonate-bearing meteoric water.

(2) Increase in the angularity of the edges of the feldspars and rock fragment grains

resulting from dissolution.


73

(3) Quartz and rock fragments become progressively more fractured, probably as a

result of matrix dissolution increasing the pressure on grain contacts.

(4) Increase in overall iron staining and increase in thickness of iron oxides around

the edges of the quartz grains and rock fragments with increase in weathering (Fig.

4.12).

(5) Increase in chlorite content and diagenetic alteration of chlorite enhances the

weatherability of rock fragments.

(6) Biotite becomes progressively degraded to become very pale hydrobiotite.

(7) Slight decrease in grain to grain contact and perfection of packing compaction

resulting from matrix and cement dissolution.

(8) Increase in the amount of matrix with increase in weathering (Fig. 4.13).

(9) Decrease in the amount of chert with increase in matrix (Fig. 4.14), due to break-

d o w n of silty volcanic chert grains.

These changes are a combination of physical and chemical weathering processes. The

main chemical changes are the transformation of the cement and matrix, primarily

siderite to chlorite and matrix clay to sericite, and the overall increase in the iron oxide

content. M a n y of the rock fragments are already quite milky in thin section in the fresh

rock and their subsequent alteration is only slight. O n the other hand the physical

changes are more widespread and cover the fracturing of the grains and the overall

decrease in grain to grain contacts. It is considered that, resulting from grain and

matrix dissolution, both the physical and chemical weathering processes play a very

significant effect on the mechanical properties of rock. The clear exception is that of

the increase in iron oxide cement, however, this feature only affects the surface rocks

and the vast majority are unaffected by significant iron staining. It has been observed

in the field between Clifton and Coalcliff that there are two distinct types of shale in

the W o m b a r r a Shale. Both have an overall greenish grey colour and appear in hand

specimen to be of similar composition. O n e type is massively bedded with a hackly


74

fracture and is interbedded with sandstone units. The hackly fracture only occurs on

exposed and weathered surfaces. The other type of shale shows strong bedding and is

characteristically laminated and fissile. Slightly weathered specimens crumble easily and

flakes readily fall off. This distinction in shale types is most important because it has

a marked effect on the stability of coastal cliffs. The different textures could be due

to the mechanism of sedimentation. It is believed that the massive shale consists of

randomly orientated flocculated clays and fine-grained quartz. The fissile shale consists

of dispersed clay, which have ideal parallel orientation, but the parallel orientation has

been disrupted in places by coarse quartz grains and rock fragments. These coarse

grains would have the effect of increasing the fracture permeability of the shale and so

this fissile shale is much more easily weathered.

4.5 MINERAL IDENTIFICATION USING X-RAY DIFFRACTION

4.5.1 INTRODUCTION

X-ray diffraction is a valuable tool in determining the mineralogy of sedimentary rocks.

This is especially important for claystone, shale and weathered rocks where petrographic

methods are of limited utility (Lewis and McConchie, 1994).

The basic concept in X-ray diffraction is that a beam of X-rays is diffracted from the

lattice planes of any crystalline substance. This phenomenon is similar to the reflection

of white light from a plane mirror. It should be noted that non-crystalline (amorphous)

substances such as glass do not produce a diffraction pattern.

4.5.2 AIM OF STUDY

Knowledge of clay mineralogy is very important in engineering geology. These minerals

are common in soil, claystone, shale, sandstone and limestone. One member of this

family is smectite (montmorillonite) which is known as a swelling or expanding clay,


75

because it expands w h e n saturated with moisture, and can cause serious problems such

as:

(1) swelling clays can contribute to instability of cliffs and rock slopes especially in

the region above the groundwater table where the rocks are subjected to alternating

wet and dry conditions as a function of rainfall and infiltration;

(2) swelling clays can cause mass movement because expansion of clays occurs during

or after periods of heavy rainfall in hillslope materials (e.g. the talus deposits);

(3) expanding clays can cause failure of open mine faces or failure of parts of

underground workings; and

(4) swelling clays can cause uplift of foundations, and damage to tunnels, railway lines

and roads.

The aim of this study is to identify the clay species present in the soils and weathered

materials and to compare them with clay minerals from weathered and fresh rocks and

subsoils. This should enable a determination of the effects of weathering in regard to

slope stability.

4.5.3 METHOD OF STUDY


Method of study can be divided into six sections as follows.

4.5.3.1 SAMPLE COLLECTION


M o r e than one hundred samples were collected for this study. Samples of weathered

rocks, and surface soils were collected from the northern Illawarra district between

Coledale and Stanwell Park Station. The samples included weathered sandstone

interbedded with coal in the upper Illawarra Coal Measures, weathered shale interbedded

with the Coal Cliff Sandstone, fresh and moderately weathered Wombarra Shale, highly

weathered Scarborough Sandstone, weathered shale interbedded with the Scarborough

Sandstone, fresh and moderately weathered Stanwell Park Claystone, highly weathered
76

Bulgo Sandstone and weathered shale interbedded with the Bulgo Sandstone. These

samples were studied using X-ray diffraction to determine the variability in clay

mineralogy, which may accompany an increase in weathering, and the associated

changes of mechanical properties of the rocks. For comparison, fresh samples were

selected from borehole IL55 between the depths of 431.20 m in the Wombarra Shale

to 347.7 m in the Stanwell Park Claystone. Weathered Bulgo Sandstone and

interbedded shale were collected from Stanwell Park Station. Weathered Stanwell Park

Claystone was selected from north of Coalcliff Station and beside the Harbour Fault

between Coalcliff and Clifton. Weathered Wombarra Shale was sampled from beside

the Jetty Fault, north of the Jetty Fault, the road down to the old Coalcliff tunnel po

and between Wombarra Station and Coledale Station. Weathered shale interbedded with

Coal Cliff Sandstone was collected from the road down to the old Coalcliff tunnel

portal. Weathered sandstone interbedded with coal in the upper niawarra Coal Measures

was selected from the base of the Moronga Park slump at Clifton. Talus samples were

collected from all the studied landslide areas (head, crown and toe).

4.5.3.3 WHOLE ROCK ANALYSIS


The most basic application of X-ray diffraction for mineral identification is in the

analysis of whole rock samples. For good results, grain size must be reduced to a

diameter of 5-10 microns. Care must be taken in particle size reduction because

crystalline grains may be damaged during grinding.

For sample preparation, first the sample must be ground to a fine powder. This can be

done quite simply with a mortar and pestle or mechanical grinder. In this study both

techniques were used. First, the rock samples were reduced to fragments less than

10 mm in diameter with the aid of geological hammer. Approximately 50 grams of this

reduced fraction was crushed in a mortar and pestle over a period 30-45 minutes. The

mechanical crusher took between 5 and 15 seconds depending upon the hardness of the
77

sample, usually harder and fresher samples taking 15 seconds, while the weathered and

softer samples took 4 to 10 seconds.

The crushed sample was mounted in an aluminium holder and X-rayed. A simple

qualitative analysis is required for the interpretation of the chart record (e.g. Fig. 4.15).

The 20 angle increases from left to right on the horizontal, and the intensity of the

diffracted peak is given by the vertical scale. The peaks are identified in terms of 20

which is converted to lattice spacing. For experimental purposes the position of the

diffraction peak is taken as the position of the point of greatest intensity.

As all minerals and all crystalline materials possess a unique X-ray diffraction pattern,

a comparison of diffraction patterns of the unknown mineral phases with a set of

standard patterns wUl lead to their identification. This method is very similar to human

fingerprint identification (Hardy & Tucker, 1988).

These standard patterns have been compiled by an international organisation called the

Joint Committee on Powder Diffraction Standards (JCPDS), which collects and updates

powder diffraction data. In principle, by a systematic searching of the JCPDS powder

diffraction index for minerals, it is possible to identify almost any mineral that may be

present, provided that sufficient peaks are present for that mineral (usually a minimum

of three). This operation can be carried out by computer using the JCPDS file and a

commercially available computer system.

Problem may arise with complex mineral mixtures since numerous peaks will be

produced which may lead to erroneous combinations.

4.5.3.4 CLAY MINERAL ANALYSIS

X-ray diffraction is the most efficient method for determining the presence of clay

minerals in claystone, shale (Surendra and Lovell, 1984), sandstone and limestone.

Although some clay minerals are evident in whole rock diffractograms, the most

satisfactory method is to extract and separately analyse the clay fraction (usually defined
78

as less than 4 microns). It is particularly important to do this in the case of very fine

grained and poorly crystalline clays, which are unlikely to give a recognised diffraction

pattern in a whole rock scan (Hardy and Tucker, 1988).

In general, for identification of clay minerals by X-ray diffraction a sample must be run

on the diffractometer three times. On the first run an untreated oriented sample,

prepared by suction onto a ceramic disc, is used. On the second run the sample is

glycolated to expand any expandable clays. For the third run a sample that has been

heated to 600C for 1 h to 2 h is used (on the ceramic disc). This heating causes some

clay lattices to collapse.

4.5.3.5 NORMAL OR UNTREATED SAMPLES

Such samples were run after they had come to equilibrium with the atmosphere in a

silica gel desiccator. They were run at 2 20 per minute, initially between 4 and 70

20, at scan speed of 2 per minute. For Cu radiation all (001) basal reflections for

kaolinite, chlorite, illite, montmorillonite (smectite) and mixed layer clays lie between

4.5 and 14 20. The (002) and (004) reflections of kaolinite and chlorite occur at

24.85 and 25.15 20 respectively.

4.5.3.6 GLYCOLATION

After the samples had been run in their untreated state, they were glycolated and run

again between 4 and 16 20. The object of the glycolation was to expand the

expandable clays to aid identification.

The process of glycolation is as follows: clay samples were prepared by suction system

on the ceramic discs and were placed in a vacuum desiccator, one or two drops of

ethanediol (ethylene glycol - CH2OH.CH2OH) was gently placed in the centre of each

disc. Samples were ready for X-ray analysis between 1 and 2 hours later.
79

4.5.3.7 HEATED SAMPLES

One episode of heating was carried out on each sample. Samples were heated to 600C

in a muffle furnace for one hour. After the sample was removed from the muffle

furnace it was allowed to cool, placed in a silica gel desiccator and then rerun between

4 and 16 20. This was to measure the true area of the 8.85 2 (10A) 001 illite peak

(Gillott, 1989). X-ray analysis was carried out on batches of 7 to 10 samples, with

each batch being glycolated and heated as a group.

4.6 RESULT OF X-RAY ANALYSIS

The results of the relative abundance of minerals for whole rock samples are tabulated

in Tables 4.9 to 4.21 and for fill and talus materials the data are provided in Tables

4.22 to 4.25.

Sixty four whole rock samples were used for determining the bulk mineral composition

(claystone, shale and highly weathered rocks), 26 samples were used for determining the

composition of the talus (Table 4.23 to 4.25), and 10 samples were used for determining

the composition of the fill (Table 4.22). One hundred clay samples were used to

determine the type of clay minerals in the clay-rich samples (ceramic discs).

The raw intensity of the main clay minerals kaolinite, illite and smectite

(montmorillonite) vary from trace amounts in some samples to a moderate amount in

other samples.

Kaolinite content is generally moderate. Illite is usually present in rare to trace amoun

but is moderate in sample H4 (Table 4.22) from the Harbour slip and common in

sample J9 (Table 4.23) from the Jetty Fault slip.

Smectite (montmorillonite) is commonly in rare to trace amounts in the majority of

samples. Quartz is abundant in all samples.

Feldspar has not been detected in many samples. Some samples have a little orthoclase

(rare to trace) but it is common in sample C3 (Table 4.25). Micaceous minerals are
80

generally few to rare except in samples Jl, J9 and sample VSS4 (Table 4.9; highly

weathered sandstone belonging to the Illawarra Coal Measures) where they are commo

Carbonate minerals (Ca, Mg, Ba) are usually rare to trace in the whole rock sample

Goethite is mostly rare, but it is generally moderate in fill derived from weather

interbedded lower strength rocks (Harbour slip). This mineral has not been detecte

the volcanic sandstone samples (VSS1-VSS5) from the Illawarra Coal Measures.

For this study the result of glycolation and heating of one sample is shown graphi

in Figure 4.16 and described in more detail in the smectite section.

4.6.1 KAOLINITE

In general, the determination of kaolinite group minerals by X-ray diffraction is

All (001) basal reflections for kaolinite lie at 12.28 20 (7.15A) and the (002) r

lies at 24.94 20 (3.57A). Usually higher-order reflections are too weak for recog

in samples composed of several clay minerals. Disordered (poorly crystallised) kao

shows broader and less intense basal reflections (Lindholm, 1987).

After glycolation there is no change for this mineral but after heating to 600C,

kaolinite structure collapses (peaks disappear). Kaolinite and chlorite both have

reflections in untreated samples. After heating, the kaolinite loses its crystalli

character and its 7A reflection disappears; chlorite is also affected by heating a

7A peak collapses, but its 14A peak is enhanced after heating, thus confirming the

presence of chlorite.

4.6.2 ILLITE

Illite generally has a broad (001) reflection at approximately 8.8 20 (10A) with

integral series of basal reflections including 17.7 20 (5A) and 26.75 20 (3.3A).

glycolation there is no change for this mineral but after heating the (001) reflec

become more intense.


81

4.6.3 S M E C T I T E ( M O N T M O R I L L O N I T E )

The basal reflection of smectite is variable from 6.80 to 5.89 20 (13-15A),

order basal reflections are irregular. Glycolation is accompanied by expansion

mineral. The (001) reflection increases to approximately 5.2 20 (17A) with an

series of basal reflections including 10.4 20 (8.5A) and 15.5 20 (5.7A). Afte

the (001) reflection collapses to between 9.83 and 8.84 20 (between 9 and 1

corresponding decreases for the integral series of higher-order basal reflecti

(Lindholm, 1987).

4.6.4 MIXED-LAYER CLAYS

Disordered mixed-layer clays are difficult to identify, in some instances wher

the dominant clay mineral present in a sample, there is sometimes a marked "tai

present on the low-angle side of the 10A peak. This is taken to indicate the p

of mixed-layer clay minerals. Following sample treatment with ethylene glycol,

illite peak broadens towards the high angle side. This may be interpreted as an

indication of disordered mixed-layer minerals (Lindholm, 1987).

After heating to 600C mixed-layer clays are identified by a the presence of a

peak at 8.84 20 (10A), a broad peak between 7.8 and 9 20, and illite is iden

the presence of a sharp peak at 8.84 20 (10A).

4.7 MUD ROCKS AND SANDSTONES

At the base of Moronga Park slump, highly weathered sandstone consists of main

quartz, mica and kaolinite; feldspar and carbonate are rare (Table 4.9). Weathe

shale interbedded with the Coal Cliff Sandstone (Table 4.10) consists of mainl

but mica, kaolinite and smectite are in rare to trace amounts and carbonate co

minor to rare.
82

Moderately weathered Wombarra Shale (between Coalcliff and Clifton, and beside the

Jetty Fault) consists of quartz, kaolinite, mica and smectite (Table 4.11). There is no

significant difference in composition between the fresh (Table 4.12) and moderately

weathered Wombarra Shale. Tables 4.13 and 4.14 show the results of X-ray diffraction

for moderately weathered Wombarra Shale between Coalcliff and Clifton north of Jetty

Fault and between Wombarra and Coledale. The latter consists of mainly quartz but

mica, smectite and illite are rare to trace and kaolinite is fair to trace. The former

comprises mainly quartz but kaolinite and carbonate are rare and smectite is fair to ra

Highly weathered Scarborough Sandstone (Table 4.15) comprises mainly quartz, mica,

kaolinite and smectite. Feldspars and carbonate are rare. Weathered grey shale

interbedded with the Scarborough Sandstone (Table 4.16) consists of quartz, mica,

kaolinite and smectite without feldspar.

Fresh Stanwell Park Claystone in the West Cliff (Table 4.17) consists of abundant

quartz, a moderate amount of kaolinite and mica, and a fair smectite but no feldspar or

carbonate. Fresh Stanwell Park Claystone in the north of Coalcliff Station (Table 4.18)

consists of common quartz, a moderate amount of kaolinite and illite, a fair smectite a

a fair to moderate amount of carbonate. Moderately weathered Stanwell Park Claystone

(Table 4.19) comprises fair to moderate quartz, kaolinite, carbonate, smectite, moderate

illite and rare mica.

Highly weathered Bulgo Sandstone (Table 4.20) consists mainly of quartz, mica,

kaolinite and smectite. Weathered grey shale interbedded in the Bulgo Sandstone (Table

4.21) consists of quartz, kaolinite, mica, smectite and illite.

The fresh and moderately weathered Wombarra Shale samples have a similar

composition. Both of them contain significant amounts of all three clay mineral types

and no significant change in the relative proportions of these minerals takes place wit

an increase in weathering. Hence, physical weathering processes and not chemical

weathering appear to be dominant. On the other hand, the Wombarra Shale has a
83

different composition to the Stanwell Park Claystone and hence, it has different

mechanical properties to the Stanwell Park Claystone (to be discussed in chapter 7).

Fresh Stanwell Park Claystone compared to weathered Stanwell Park Claystone has more

quartz and less clay minerals. The samples appear to show a change in the clay

mineralogy with an increase in chemical weathering. Grey shale interbedded with the

Bulgo Sandstone a moderate content of mixed-layer clays and smectite which generally

makes the Bulgo Sandstone relatively unstable.

4.8 TALUS

Generally the talus is a thin, surficial cover of unconsolidated material commonly

between 0.5 and 2 m in depth, but in some places attaining depths of up to 40 m. It

consists of blocks, up to several metres in size, of Hawkesbury Sandstone and Narrabeen

Group sandstone in a clayey matrix which may be iron stained or leached. The relative

proportion and size of the sandstone blocks vary considerably from area to area.

4.8.1 COMPOSITION

As mentioned earlier the talus consists of blocks of sandstone in a sandy, clayey matrix.

Forty one soil samples were collected from the landslide area (head, crown and toe).

Ten samples were selected from the Harbour slump beside the Harbour Fault in

Coalcliff Harbour. Nine samples were collected from the Jetty rock slump beside the

Jetty Fault between Clifton and Coalcliff. Ten samples were selected from the Moronga

Park slump and seven samples were collected from the Clifton Hotel slump beside the

Clifton Hotel.

Generally talus consists of quartz, mica, kaolinite, smectite, illite and goethite (Tables

4.22 to 4.25). Kaolinite is the dominant clay mineral in all the talus. Mixed-layer

clays are important in the talus material in whole rock samples. Commonly the more

unstable talus samples have a higher proportion of mixed-layer clays and smectite than
84

the more stable ones. This results from the lower shearing strength of smectite, mixed-

layer clay and illite-rich talus compared to the kaolinite-rich talus.

4.9 INTERPRETATION OF THESE RESULTS


Kaolinite is the dominant clay mineral in talus and rock samples. This mineral may be

either a result of direct derivation from the Narrabeen Group (Wombarra Shale, Stanwell

Park Claystone) or be the result of weathering of talus during past periods of higher

rainfall. The present average yearly rainfall is about 1700 mm and the mean daily

temperature is 17.5C; even these condition would produce complete weathering of the

clay fraction to produce kaolinite over a period of a few thousand years. Also, several

talus samples showed that they have a large amount of kaolinite and less illite and

smectite which may indicate that they represent more mature (weathered) talus.

4.10 CLAY MINERAL STRUCTURE AND SLOPE STABILITY


The relationship between slope stability and clay mineralogy can be explained by the

structure of the clay minerals, which influences the behaviour of any soil containing

these clays. In particular the shear strength and deformation of soils depends on the

type and content of clay present in that soil.

Clay minerals consist of different stacking arrangements of tetrahedral sheets and

octahedral layers. Kaolinite has a stacking of one tetrahedral sheet to one octahedral

sheet and has no ionic charge deficiency. Formula units of kaolinite have relatively

strong hydrogen bonding, unlike other clay layers which have weaker oxygen bonds.

As a result, kaolinite would be expected to have the strongest cohesiveness and greatest

stability of all clay minerals. Another result of the strong hydrogen bonding is that

water molecules have difficulty penetrating between the layers. In haUoysite, a mineral

of similar structure to kaolinite, water can penetrate between the tetrahedral and the
85

octahedral layers and, as a result, this mineral has a m u c h lower cohesiveness (Grim,

1968; Gillott, 1989). However halloysite has not been detected during the present study.

Illite consists of two tetrahedral sheets, with some aluminium substitution for silicon,

sandwiching an octahedral sheet stacked so that oxygen layers of each unit are adjacent

to oxygen layers of the neighbouring units, resulting in a very weak oxygen-oxygen

bond between sheets. This gives a good basal cleavage and hence a weaker shear

strength or cohesiveness than kaolinite.

Structurally, smectite is similar to illite except that no substitution of silicon by

aluminium occurs in the tetrahedral sheet. If all the octahedral positions in smectite are

occupied the mineral is termed trioctahedral, and if only two-third of the sites are

occupied the mineral is dioctahedral. The second type is most c o m m o n in soil, and

water can enter between the layers and cause lattice expansion.

The ability to absorb water is influenced by the exchangeable cations. Thus sodium

montmorillonite (smectite) can absorb more water than calcium montmorillonite (Franklin

& Dusseault, 1989), and is more sensitive to water with a resulting greater loss in

cohesiveness.

The whole rock sodium to calcium and magnesium ratio m a y be relatively moderate for

sandstone in the Narrabeen Group and Hawkesbury Sandstone (Bowman, 1972). Thus

sodium montmoriUonite would be expected to form on weathering in the niawarra area

more than calcium montmorillonite. Therefore, it will be a cause of slope instability in

the area.

4.11 DISCUSSION

Petrology and weathering of rocks are two very important aspects from an engineering

geological point of view. A s mechanical properties of rocks are sensitive to their

petrology and weathering, evaluation of both petrology and degree of weathering of the

rocks is important for different study purposes, such as the estimation of long-term
86

stability of the slopes. The relationship between mineralogical and mechanical properties

of rocks and their degree of weathering has been studied by many research workers.

For example Ramana and Gogte (1982) proposed the percentage of decomposition as

an index of the weatherability of rocks, and Gunsalus and Kulhawy (1984) obtained a

positive correlation between shear strength parameters and quartz content.

Petrological studies reported in this chapter indicate that the Narrabeen Group sandston

in the northern Illawarra between Scarborough and Stanwell Park consists principally of

detrital quartz grains, lithic fragments (including chert) with detrital feldspar grains.

The well-rounded quartz grains probably represent second cycle sediments and were

derived from earlier deposited sandstones. The poorly rounded to angular quartz clasts

were derived from volcanic sources. Volcanic rock fragments mostly comprise chert

grains, probably derived from silicified silicic volcanic rocks and tuff.

All these studies show that the content of detrital quartz grains generally increases

towards the top of the group and that the detrital lithic fragment content generally

decreases. Some fresh samples of Bulgo Sandstone (lower part) and Scarborough

Sandstone contains more than 50% quartz, while some samples of Coal Cliff Sandstone

only contain about 30% quartz. Therefore, the Bulgo Sandstone and Scarborough

Sandstone are stronger than the Coal Cliff Sandstone.

The percentage of rock fragments in the Coal Cliff Sandstone is more than the

percentage of rock fragment in Bulgo Sandstone and Scarborough Sandstone.

Consequently the sandstones in the Coal Cliff Sandstone are easier targets than the

Bulgo Sandstone and Scarborough Sandstone for long-term weathering. This is most

important when it is considered that rock fragments are mostly chert and volcanic

detritus in the Narrabeen Group.

The middle of the Bulgo Sandstone consists of green sandstone and shale. The

sandstones comprises predominantly volcanic rock fragments which are altered to

chlorite and iron oxides. These sandstones have a higher proportion of rock fragments
87

than quartz and, therefore, they are easy targets for long-term weathering and

decomposition.

The majority of the sandstone samples contain carbonate cement. The carbonate

cements include calcite, siderite, and perhaps ferroan calcite, dolomite and ankerite. In

some thin sections, partially dissolved carbonate cements were recognised. The

dissolution has created secondary pore spaces. For short-term weathering, the quantity

of carbonate cement is a useful parameter for determining the stability of the rock.

During long-term weathering, dissolved carbonate cement causes more pore space in the

rock, and consequently secondary porosity and permeability increases. The resultant

increase in water flow into the rock mass can affect its weathering. Most of the effects

of weathering are concentrated along cliffs and in the soft underlying shale beds. For

example, fretting and weathering of the Wombarra Shale have undermined the

Scarborough Sandstone. Also fretting and weathering of the Stanwell Park Claystone

have undermined the Bulgo Sandstone. This causes large blocks of the overlying

sandstone to break off, either sliding or toppling.

Results of X-ray diffraction analyses indicate that the rocks and talus materials typically

contain quartz, iron oxides, kaolinite, illite, smectite and expanded-lattice mixed-layer

clay minerals. Quartz and kaolinite are abundant in most rocks and talus materials in

the northern Illawarra.

Clay minerals consists of kaolinite with less illite and smectite (montmorillonite). Most

clay samples mainly contain kaolinite which is poorly crystalline. On the other hand,

some of the clay samples mainly consist of mixed-layer clays (illite and smectite).

These clays cause swelling and shrinkage near the surface. They lead to decrease in

the shearing resistance of the talus and surface rock materials. Decreased shear strength

increases the susceptibility to mass movement.

X-ray analyses proved that in the transition from fresh rock samples to weathered rock

samples the amount of primary minerals such as feldspars decrease while that of
88

secondary minerals (i.e clay minerals) increases. As the proportion of secondary

minerals such as clay minerals and iron oxides increases, the strength of the rock

decreases while the porosity increases.

The results revealed that the petrology and weathering phenomenon affect both physical

and mechanical behaviours of the rocks. The physical and mechanical properties of

rocks change with change in mineralogy and the degree of weathering.

4.12 RELATIONSHIP BETWEEN PETROLOGY, SOURCE AND SLOPE


STABILITY
The lithology of the source area exerts control on the abundance of sand-size rock

fragments released into the transport mill. The tectonic setting, which ultimately

controls source-rock lithology, has an influence also on rock-fragment abundance.

Dickinson and Suczek (1979) and Dickinson (1985) suggested in their provenance

diagrams that rock fragments are more abundant in sediments derived from magmatic

arcs than in sediments derived from recycled orogenic or continental-block provenance.

The Narrabeen Group was derived from the New England Fold Belt to the north and

consists predominantly volcanic detritus. The volcanic detritus is present in both the

sandstone and shale units either in form of detrital grains of volcanic rock or as fine

volcanic ash. During post-depositional alteration and diagenesis the original volcanic

glass in the ash and matrix of larger grains has devitrified to produce smectite clays.

These clays not only cause swelling and shrinkage near the surface as a response to

wetting and drying, but also reduce the permeability of the near surface rock mass.

This latter factor increases the aqueous pore pressures and hence increases the likeliho

of surficial mass movement of both the rock mass and the adjacent talus deposits.

Based on the X-ray diffraction analysis, carbonates are mostly rare in the talus deposit

(Tables 4.23 to 4.25). The natural reduction in the carbonate due to dissolution on

weathering is a factor in the talus slope instability in the Illawarra area. Hawkins et
89

al (1988) have discussed the similar importance of calcite content in the stability of

Fuller's Earth in the UK. Hawkins and McDonald (1992) pointed out the importance

of the decalcification which effects the instability in both natural and engineered slopes.

Trotter (1993) reported earthflows which developed in regolith overlying Tertiary

calcareous mudrock of marine origin in the North Island, New Zealand. He showed that

slope stability is reduced by natural reduction in the calcite content as a result of

exposure and weathering.


90
91

CHAPTER 5
STRUCTURAL GEOLOGY IN THE SLIP AREA

5.1 INTRODUCTION
It is well known that the properties of intact rock units are often relatively unimportant

as controls for slope stability in rock, and that failures are governed by structural

discontinuities such as bedding planes, joints and faults in the rock mass (Hoek, 1971).

The frequency, orientation and inclination of these discontinuities, as well as their

strength characteristics, are important factors to be considered in any rock slope analysis.

They also control the flow of groundwater which has a very important influence on

slope stability. Hence, the need to locate and establish the orientation and strength

properties of the critical discontinuities in a rock mass is obvious.

5.2 STRUCTURAL FACTORS WHICH ARE IMPORTANT IN SLOPE


STABILITY
Theoretical studies and practical experience of rock slope problems suggest that the

structural factors which are important in slope stability are as follows:

(1) The dip of joints, faults and bedding planes. This inclination plays a major role

in defining the stability of a slope.

(2) The strike of structural features. Obviously the relationship between the strike

directions of important structural features and the orientation of a slope face will

define the freedom of movement of potentially unstable blocks. The most

dangerous conditions occur when an unfavourably inclined structural feature has

a strike parallel to the slope face such that the entire face is free to slide.

(3) The number of structural features or sets of such features. The behaviour of a

slope containing a single set of bedding planes is likely to differ from that of a
92

slope containing additional sets of intersecting joints. Not only will the entire rock

mass have greater freedom to deform in the latter case but the flow of

groundwater, which plays a major role in controlling the stability of a slope, will

be markedly different in the two cases.

Spacing. Spacing is defined as the average distance between adjacent

discontinuities in a set, measured normal to the discontinuity plane. Descriptions

can be used such as widely or closely spaced joints, thickly or thinly bedded. For

engineering purposes, these spacings should always be defined. The block size is

related to joint and bedding spacing in that it is governed by, and approximately

equal to, the average spacing for all sets in the jointing-bedding system. Intensity

of jointing is the inverse of spacing, i.e. the number of joints per metre. As an

alternative parameter to spacing it has the practical disadvantage that large number

must be counted when joints are closely spaced (e.g. 1 cm, or less).

The continuity or persistence of structural features. Structural features are often

modelled as disk shaped, or as terminating in straight edges where they meet

other, nOn-parallel joints (Merritt and Baecher, 1981). In stability calculations for

any slide area, the persistence of joints has a great effect on the shear strength,

and an engineer needs to know the total surface area over which sliding could

occur. Faults, because of their mode of formation, can be inferred as 100%

persistent and often are plotted for many kilometres on geological maps. Most

bedding joints are also highly persistent usually on a regional scale.

The surface properties and infilling material in joints and faults. Surface

characteristics such as roughness and the nature and thickness of any infilling

material which may be present in a structural discontinuity have been shown to

have a major influence upon the shear strength characteristics of such

discontinuities. The permeability of these discontinuities is also significantly

influenced by the presence of filling materials. Filling materials vary greatly in


93

their mechanical characteristics, from very soft to very hard and strong. Materials

of extremely low strength, that call for precautions w h e n encountered, include clay,

platy and very soft minerals, such as smectite, illite, chlorite, graphite and talc.

Clays are soft and soapy in texture and usually have a high water content. W h e n

they have a bearing on rock mass stability they should be further characterised

by X-ray diffraction or optical mineralogical testing and by plasticity and water

content measurements. The clay mylonites found in shales consist of finely

ground and highly plastic shale created by shearing. Even w h e n only 1 to 2 m m

thick and dipping at a few degrees, they have been k n o w n to cause extensive slope

failures. Fillings of intermediate strength include those of sandy consistency, such

as crushed or brecciated hard or moderately hard rock, lightly altered wall rock,

or veins of calcite w h e n weaker than the surrounding wall rock. Strong fillings

such as vein quartz, calcite and limonite can heal and recement a joint, which m a y

become as strong as the surrounding rock or even stronger than it if the latter

becomes weathered.

Most joints have no filling and are, therefore, neither strengthened nor weakened

byfillingmaterials. Their unaltered surfaces are in contact. Stained joints can be

included in the "no filling" category. Staining is important as an indicator that a

joint has conducted groundwater, but usually has little or no influence on strength

or other mechanical properties.

Aperture. The aperture of a joint (also known as its openness or separation) is

the m e a n distance separating the two intact joint walls. Note that aperture includes

the thickness of anyfillingthat m a y be present. A joint m a y be termed open or

tight according to whether its aperture is large or small. The aperture is usually

greatest for near-surface joints, as a result of rebound and stress release, and joints

become tighter as the depth increases. Apertures are usually just a few microns
94

wide, except where the rock has been loosened by near-surface weathering or

blasting, or dissolved by water flowing through the joints.

Therefore, it is important that adequate provision is made for the collection of structur

information (Bell and Pettinga, 1984) and this means that a geologist who has some idea

of what structural features are important in controlling slope stability must be allocate

to the task of structural mapping.

5.3 FAULTS IN THE SLIP AREA


The study area has been subject to extensive geological activity in the past. Several

major faults and dykes are present which have resulted in the formation of prominent

topographical features such as the creek beside Clifton Fault. The majority of faults in

the study area possess strike maxima at 005 and 110 (Bowman, 1974).

The major faults which are identifiable on the surface are named the Scarborough Fault,

Clifton Fault, Jetty Fault and Harbour Fault (Fig. 5.1). Several other fault zones were

identified in mine workings but do not extend to the surface. The largest is the

Coalcliff Fault. This fault at the level of the Bulli seam has a dip to south of about

70 from horizontal. Some other smaller faults, although present in the Bulli seam,

could not be located on the surface (cf. Fig. 2.11). This may either be due to the

blanketing effect of the talus mantle or simply that they do not project to the surface.

Two small un-named faults occur in the Bulgo Sandstone (Fig. 5.2) and Stanwell Park

Claystone (Fig. 5.3) between Clifton and Coalcliff.

It appears that the faults with east and southeast trends observed in the mine working

have little or no displacement at the level of the Hawkesbury Sandstone. The north-

trending faults however show displacement in both the Bulli Coal seam and the

Hawkesbury Sandstone (Bowman, 1974). North-trending faults remained active at least

until after the deposition of the Hawkesbury Sandstone, whereas the east-trending faults
95

had become inactive or were only weakly active by this stage (Bowman, 1974;

Adamson, 1974).

East-west striking (110) faults are concentrated in the coastal area between Clifton and

Coalcliff. This group includes the Scarborough, Clifton, Jetty and Harbour Faults, all

of which appear to have steeply dipping fault planes. These faults are most readily

observed because of the rugged coastal outcrop in this locality which is almost normal

to the fault trend.

5.3.1 HARBOUR FAULT


The Harbour Fault was named by Hanlon (1953), as were the other faults in the study

area. This fault is seen to dislocate the lower section of the Narrabeen Group in the

cliffs adjacent to the old Coalcliff Harbour, south of Coalcliff Beach (Fig. 5.4).

The fault has cut the rock platform with a readily distinguished lineament. A

downthrow to the south of 20 m was measured by the author in the coastal platform

at the top of the Coal Cliff Sandstone. T o the west of the Lawrence Hargrave Drive,

erosion has revealed the fault plane surface, which dips at about 70 to the south and

strikes east-west.

5.3.2 JETTY FAULT


The Jetty Fault crops out in the coastal cliff a short distance north of the old Coalcliff

adit. The dip of the fault plane is 70 north on Lawrence Hargrave Drive and 45 north

in the lower coastal cliff north of the old Coalcliff adit; its strike is east-west (Figs 5.5a,

5.5b). This fault is particularly well exposed and Hanlon described a decrease in throw

with stratigraphic ascent. A displacement of about 8 m at the Bulli seam level

decreases to about 7.5 m at the top of the Coal Cliff Sandstone, about 2.8 m at the

level of the Otford Sandstone M e m b e r and about 1.1 m at the base of the Scarborough

Sandstone (Fig. 2.11).


96

5.3.3 CLIFTON FAULT

The Clifton Fault crosses the coast just north of Clifton. The strike of the fault is east

west, and it is downthrown to the north. The dip of the fault plane is nearly vertical

in the coastal cliff exposure (Fig. 5.6). It is marked by a prominent, straight creek and

an abrupt termination of the coastal platform. The Bulli Coal crops out approximately

at sea level on the northern side of the fault. The fault has a displacement of up to

60 m in the study area.

5.3.4 SCARBOROUGH FAULT

The Scarborough Fault crosses the coast just north of Wombarra (Fig. 5.7). The vertical

displacement is as given 55 m on old plans from the southern workings of CoalcHff

Colliery. Harper (1915) gave the dip of the fault plane as 15 north but Hanlon (1953)

measured the dip of the fault plane to be between 40 and 60 north. The dip of the

fault is certainly steeper in the Triassic beds in the area. There is evidence of splittin

of the fault in both easterly and westerly directions. It is possible the main Scarborough

Fault swings away to the southeast.

5.4 JOINTS IN THE SLIP AREA

Joints are discontinuities in the rock mass along which there has been no relative

movement. When there has been movement along the structure then it becomes a fault.

For this reason, in many cases, it is difficult to identify a small fault when no

lithological variations are present permitting determination of movement.

The most significant structural feature in the escarpment is jointing. On the local scale

of the escarpment, folding of the strata is generally negligible and the bedding is very

close to horizontal a dip of 2-5 to the northwest. Faulting is common and, although

fault zones influence the groundwater, it generally plays a very similar role to jointing
97

w h e n considering discontinuities as an influence on the escarpment's stability. During

the measurement of joints for this study, examination of aerial photographs revealed the

predominant pattern which is closely related to the pattern of faults exposed in the

underground workings.

The most prominent joint set exposed at the surface is close to vertical with a north-

south trend or a trend a few degrees east of this direction. This joint direction is

subparallel to the main trend of cliffs in the study area. Hence the prominent sandstone

beds generally outcrop as vertical or near vertical faces.

The frequency of joint spacing in the various rock units varies. Joint spacing has been

measured in this study in an attempt to predict h o w this property m a y affect rock slope

stability in the niawarra area.

5.4.1 JOINTS IN COAL

Joints in coal have been recognised and utilised by coal miners for centuries in the

layout of mines and in the mining of coal beds (Kendall and Briggs, 1933). Coal beds

of lignitic to bituminous rank are usually cut by an orthogonal system of two mutually

perpendicular joint sets; a prominent set termed cleats or face cleats and a secondary

set called cross-cleats or butt cleats. Butt cleats or butt joints commonly break away

at the face cleats or face joints which produce the larger and more continuous faces

seen in the mine. Joints in coal vary in spacing from 1 millimetre or less to several

centimetres. Widely spaced joints divide the bed into rectangular blocks, giving rise to

the miner's term "block coal". Where the coal bed is relatively homogeneous, jointing

is closely spaced. Also joint spacing varies with the amount of weathering. Well-

weathered coals show more joints than do fresh exposures. Coals are distinct structural

rock units because they display joint systems differing in orientation and frequency

from those of other rock types. In addition, bright bands within the coal beds show
98

more prominent and, in places, differently oriented joints than do dull bands. Bright

bands thus act as minute structural rock units within the coal bed.

5.4.2 JOINTS IN THE NARRABEEN GROUP AND HAWKESBURY

SANDSTONE
B o w m a n (1974) identified joint distributions in the Hawkesbury Sandstone with m a x i m a

at 005 and 105. W e a k maxima of 055 and 155 are also present in the niawarra.

These data are based on air-photograph tracing of exposures of the Hawkesbury

Sandstone along the main scarp. The distinct vertical faces in the Hawkesbury cliff line

also highlight an obvious subvertical joint group. Joint spacing has been observed to

be generally from 2-5 m .

The author has measured joints at specific critical locations to determine if localised

joints patterns are important. Figure 5.8 shows where joint measurements were taken.

The results are shown as computer plotted joint rosettes in Figures 5.9 to 5.14. All

joints rosettes are based on an interval of 10.

Joints were studied in the Bulgo Sandstone (BSs), Stanwell Park Claystone (SPC),

Scarborough Sandstone (SSs), Wombarra Shale ( W S h ) and Coal Cliff Sandstone (CSs).

The resulting strike maxima for the Bulgo Sandstone (localities 1, 2, 3) are 025, 045

and 105 (Fig. 5.9); for Stanwell Park Claystone (localities 4, 5) they are 025, 050 and

105 (Fig. 5.10); for the Scarborough Sandstone (localities 6, 7, 8, 9) they are 015,

035, 135 and 165 (Fig. 5.11); for the Wombarra Shale (9, 10) they are 015, 115 and

165 (Fig. 5.12); and for the Coal Cliff Sandstone (localities 10, 11, 14) they are 015,

025 045 and 145 (Fig. 5.13) and for localities 12 and 13 they are 005, 035, 065 and

135 (Fig. 5.14). A s mentioned above the most prominent joint set exposed at the

surface has a nearly north-south trend (005 to 025).

The pattern of faults is similar in orientation to one of the joint sets (105 to 115).

They appear to be tensional features which occurred as a stress relief process in the
99

Sydney Basin strata along convenient zones of weakness. The northwest-southeast set

is subparallel to the zone of tension along the Woronora Anticline an important

synsedimentary flexure within the Sydney Basin (Fig. 2.8). Several prominent joint sets

also exist and strongly influence the surface features. Jointing appears to be influenced

by the regional stress regime and localised destressing along the escarpment. The

principal north-south joints are likely to be tensional features resulting from tectonic

stress relief induced by consolidation within an east-west compressive stress field that

affected the area during the Permian and Early Mesozoic. They are also parallel to the

rift opening of the Tasman Sea in the Cretaceous (east-west tension).

Localised destressing is possible near the exposed face of the escarpment. In the

Illawarra area the horizontal in situ stress is two to three times greater than the vertical

stress (Walton et al, 1990) and thus a large degree of destressing must occur at the

escarpment face. This, combined with weathering, would result in increased jointing

closer to the outer exposed face of the escarpment. Vertical permeability (fracture

porosity) is, therefore, expected to be greatest near the outer face of the escarpment.

Jointing in the Scarborough Sandstone is typically widely spaced (1-4 m) whereas in the

interbedded sandstone and siltstone of the Bulgo Sandstone it usually shows a 0.5-

1.5 m spacing and in the intervening Stanwell Park Claystone the joints are closely

spaced (0.1-0.5 m). Joint spacing in the Coal Cliff Sandstone is usually 0.6-2 m and

in the intervening Wombarra Shale the joints are 0.2-0.6 m apart. Many of the joints

on the escarpment and coastline are filled with calcite and/or clay.

A comparison with Bowman's (1974) joint pattern shows a reasonably good match

between the local and regional joint sets (cf. Fig. 2.12, Table 5.1).

The resulting strike maxima at 005 for the Coal Cliff Sandstone (Fig. 5.14) shows a

good fit with the 005 be (Fig. 2.12) regional group of Bowman (1974). Strike maxima

at 105 for the Bulgo Sandstone (BSs) and Stanwell Park Claystone (SPC) show a good

fit with the 105 ac (Fig. 2.12) regional group of Bowman (1974). Strike maxima at
100

045 for the Bulgo Sandstone (BSs), at 050 for the Stanwell Park Claystone (SPC), and

at 045 and 065 for the Coal Cliff Sandstone (Fig. 5.14) are close to the 055 ac (Fig

2.12) local folding of Bowman (1974). Strike maxima at 165 for the Scarborough, at

145 for the Coal Cliff Sandstones (SSs, CSs) and at 165 for the Wombarra Shale

(WSh) are close to the 155 be (Fig. 2.12) local folding of Bowman (1974). Strike

maxima at 115 for Wombarra Shale (WSh) is very close to 105 ac (Fig. 2.12) regional

group of Bowman (1974). Strike maxima at 015 for Wombarra Shale (WSh) and

Scarborough Sandstone (SSs) are very close to 005 be regional group of Bowman

(1974). Total joint directions for Narrabeen Group show two prominent joint sets,

north-northeast and east-southeast in the study area (Fig. 5.15). A comparison with

Bowman's (1974) joint pattern (Fig. 2.12) shows a reasonably good match between the

local and regional joint sets. The resulting strike maxima at 105 in the study area

shows a good fit with the 105 Bowman's (1974) joint pattern (Fig. 2.12). The small

differences are probably due to local variations in the regional pattern. Strike maxima

at 035 for the Scarborough and Coal Cliff Sandstones (Figs 5.11, 5.14) and at 025 for

the Bulgo Sandstone (Fig. 5.9) have probably been caused by stress relief. The strike

maxima at 015 for the Wombarra Shale has a great effect on slope stability between

Clifton and Coalcliff (Fig. 5.16).

Considering all the localities mentioned above, it appears that the regional joint set

most important and the development of local sets is dependent on the closeness to local

faults or folds. Local stress relief probably caused some joint sets to form. The

resulting strike maxima for the lower Narrabeen Group show that the most prominent

joint set exposed at the surface, with a direction between 005 and 025, has a

significant effect on cliff orientation and slope stability in the study area.
101
5.5 THE IMPORTANCE OF FAULTS AND OTHER THROUGH-GOING
GEOLOGIC STRUCTURES
The importance of faults and through-going structures may sometimes be forgotten

because of the enormous amount of work and expense that may be involved in detailed

joint surveys and in the plotting and analysis of these data. There are relatively few

instances in which the joint orientation data turn out to be more important than

knowledge of the position, orientation, and strength characteristics of the major through-

going structures. A sketch summarising many of the reasons for the increased

significance of the through-going structures appears as Figure 5.17.

Figure 5.17a shows a rock mass with discontinuous and/or irregular rock joints; Figure

5.17b shows the same rock mass after a shearing displacement of natural origin has

occurred along one of the pre-existing sets of joints. The effects of the shearing

displacement are as follows:

(1) Continuity is increased. Therefore, the area of influence of the structure is

increased and the cohesive component of the strength is decreased.

(2) Irregularities are decreased. Therefore, one or both of the shear strength

parameters and c may be reduced and the strength parameters may approach

those of the residual shear strength.

(3) Permeability is altered. The increase or decrease in permeability can be complex.

If the change in permeability is sufficient to allow significant pore water pressures

to develop, a decrease in the shearing resistance along the fault will then be the

consequence.

(4) Weathering and alteration are common along faults. The new weathering products

are frequently clay or other silicate minerals, e.g. chlorite. Therefore, reduced

angles of shearing resistance are common.

Figure 5.18 is a sketch of a typical cross-section of a fault. The fault has a central

zone of crushed and sheared rock called fault breccia (a), flanked by fine-grained, often
102

clay-rich, fault gouge on either side (b) and with striated and slickensided surfaces foun

on the bedrock surface (c). The zone of rock (d) adjacent to either side of the fault is

likely to be more highly fractured than the surrounding country rock (e). This sequence

of materials can be referred to as the typical composite fault.

There are many variations to this sequence. For example, the breccia may be missing,

the breccia and gouge may be missing, the fractured rock may be missing and any or

all of these layers may have been re-cemented. In addition, weathering often extends

appreciably deeper along fault zones, due to the more intense jointing and alteration is

common along faults due to groundwater movement. The weathering and alteration can

superimpose additional zones of materials with different physical properties.

The most significant engineering properties of the zones in the composite fault are also

indicated on Figure 5.18. These include the low shear strength of the gouge-rock

contact.

Small faults may have little or no influence on the slope stability. Other faults or

combinations of faults can be the most significant geological factors in the analysis and

prediction of the slope stability problem. The orientation of the structure relative to th

slope is usually critical.

5.6 JOINTING AND TECTONIC FRACTURING OF ROCK


In any particular cliff section joints play a major role in the stability of the cliff,

although the overall effect is not such that the cliff is joint controlled. The cliff-line

tends to represent the surface of least resistance to block movement. The movement

of joint-bound blocks is aided by the orientations of the joints. Therefore, the

orientation and number of joint sets and the number of joints (spacing) all affect cliff

stability. The joints are often vertical or subvertical and so provide the unfavourable

orientation necessary for failure. They acts as discontinuities unable to sustain tension

and, combined with other factors, permit failure (Fig. 5.19). Any blocks which are
103

undercut by the weathering will very easily fall off under the pull of gravity (Fig. 5.20).

S o m e precipitating action is needed to remove the supporting material from the base of

the sandstone cliffs. Lubrication of the joint planes is also helpful in starting the rock

movement. O p e n joints at the crest of slopes act as drainage channels and very largely

determine the course of water in critical locations (Fig. 5.21). Root pry, due to

vegetation growing in joints, also causes them to open up and promotes failure.

The joints are often discontinuous both laterally and vertically. Thus blocks adjacent

to an unstable section m a y not necessarily become unstable at the same time. Blocks

above or adjacent to an unstable section will not necessarily be jointed in precisely the

same place or direction. So that more undermining m a y be necessary to cause failure

of some blocks than others.

Weathering is the second major factor. It is responsible not only for the overall

decrease in strength of the rock as a whole and the asperity of the joint surfaces, but

also is the primary cause of undercutting. Figure 5.22 shows strong undercutting of

sandstone cliffs (now buttressed by concrete).

5.6.1 BEDDING

Bedding appears to play only an indirect role in the stability of the cliffs. The dips are

about 5, or less, and the beds almost always dip toward the west or northwest (away

from the escarpment). For this reason the bedding planes are only the seats of block

sliding near the edge of the escarpment when the bedding plane is strongly weathered

and the surface is lubricated with groundwater.

In discussions of bedding in the lower Narrabeen Group sandstones, the effect of local

diastems is important. The diastems are often zones of dislocation between blocks in

the cliff sequence. This feature effects the stability of the cliffs because the diastem

provides a distinct break in the cliff.


104

Another feature which may prevent block sliding along suitable bedding planes is the

apparent roughness on some bedding planes. The major breaks between layers often

have ferruginous and calcareous nodules on them as well as having wavy irregularities

due to contemporaneous channelling during deposition.

Cross-stratification or cross-bedding is commonly present in granular sedimentary rocks

such as sandstones. It consists of tabular, irregularly lenticular or wedge-shape bodies

that show a pronounced laminated structure which is steeply inclined to the general

bedding. The cross-stratification planes vary from unit to unit. Most layers can be

recognised by slight changes in grain size. Some have very marked and distinct

bedding planes. The cross-stratification planes rarely affect the stability of the cliff

a whole, but some individual blocks can be removed by translational sliding along the

plane of a cross-bed. Normally the joints provide a more strongly defined zone of

weakness, but occasionally the cross-bed planes are particularly marked. Since the

general palaeocurrent trend is to the southeast (Bamberry, 1990), many of the cross-

bed planes dip toward the edge of the cliff and movement of a pyramidal block could

occur. Although this kind of movement is probably fairly rare, it could be precipitated

by gravity at the moment of failure, but the plane would have to be clay-lined and well

lubricated by water before the block could move.

It is apparent that the sandstones of the Narrabeen Group only tend to be markedly

weaker along the bedding in special cases where the cross-stratification planes are

marked by sharp fissility planes. Most of these sandstones would tend to break away

along subvertical fractures such as joints.

5.7 RELATIONSHIP BETWEEN JOINTS AND CLIFF ORIENTATION


5.7.1 STRESS RELIEF

Rock mechanics investigations (Nichols, 1980; Kulhway and O'Rourke, 1981) have

shown that many flat-lying sedimentary sequences near the earth's surface have a
105

horizontal stress equal to or greater than the vertical stress corresponding to existing

overburden. These high horizontal stresses are presumably related to previous over-

burden removed by erosion (Ferguson et al, 1981).

In the Illawarra region the horizontal in situ stress is two to three times greater than the

vertical stress (Walton et al, 1990; Hilleard, 1993) and thus a large degree of

destressing must occur at the escarpment face.

Erosion removes horizontal support from escarpment walls, which then tend to deform

internally. The mechanics of this deformational response are well understood in concept

from elastic theory, but details of this deformational response are often complex,

depending on time dependent phenomena and stratigraphic sequences. Deformational

response due to escarpment erosion in flat-lying sedimentary rocks will be enhanced by

high horizontal stress in the rocks.

Deformational response and rock discontinuities due to escarpment erosion and stress

relief in flat-lying sedimentary rocks are shown schematically in (Fig. 5.23a). The

escarpment walls are zones of extension. Rock discontinuities in the escarpment walls

reflect these deformation conditions and also the stratigraphy, particularly the strength

and stiffness of individual beds.

Erosional retreat of escarpment walls concentrates in the weaker and more deformable

beds which sometimes develop diagonal to curved shear joints, and commonly develop

shear zones at contacts with stronger stiffer beds (Fig. 5.23b). Stronger and stiffer beds

develop vertical to sub-vertical tension joints which typically do not extend across

weaker beds or bedding contacts. The spacing of tension joints in a given bed is

directly proportional to the bed stiffness. Stress relief effects diminish with distance into

the escarpment wall. The width of the relaxed zone along an escarpment wall is highly

variable.

Most deformation occurs beneath relaxed zones in the escarpment walls. Lateral

compression plus vertical load removal causes arching and buckling of beds in the
106

escarpment bottom (Fig. 5.24). Bedding planes open in strong stiff beds while fractures

develop in weaker and more deformable beds. Very gradual straining and fracturing

would be continually occurring far back from the scarp edge, but the overall effect

would be far less dramatic than in an open pit mine. Well before a particular rock

block was actually located on the edge of the scarp, due to cliff retreat over geologic

time, it would have had stress relief perpendicular to the cliff line. However, the

horizontal stresses acting parallel to the length of cliff edge often would not have been

relieved, as evidenced by joints, striking perpendicularly to the scarp, being closed.

Considering the escarpment shape (Fig. 5.23), it is clear that some situations exist where

these horizontal stresses have been relieved. However, the remaining unrelieved

horizontal stresses may act to stabilise a particular rock column by providing a lateral

shear stress inhibiting vertical movement. In other situations, these stresses may act to

induce failure.

5.8 THE RELATIONSHIP BETWEEN JOINTS AND RATES OF EROSION OF


STRATIGRAPHIC UNITS IN DIFFERENT TYPES OF EXPOSURES
The entire niawarra Escarpment has been retreating westward for many years (about 90

Ma since the opening of the Tasman Sea; Veevers et al, 1991). At present most of

escarpment south of Wombarra is protected from active marine erosion by the coastal

plain. North of Wombarra, the base of the escarpment is subject to active marine

erosion. Only positive erosion control measures prevent the Lawrence Hargrave Drive

between Coalcliff and Clifton from being completely eroded away. Therefore, in this

critical location, marine erosion is a major factor; whereas along most of the length of

the escarpment it has minor indirect effects, for example sea spray may be an active

weathering factor.
107
5.8.1 DIFFERENTIAL EROSION

Differential erosion is the result of weathering of rocks which are not uniform in

character but are softer or more soluble in some places than in other areas. The result

is usually an uneven surface with the softer rocks being removed more quickly than the

harder sequence. Differential erosion is caused mainly by the erosive processes of

rainfall and sea acting on the various lithologies in the cliffs. T h e coarser sandy

lithologies of the Scarborough and Bulgo Sandstones form the harder units which are

eroded more slowly and which jut out as cliffs. The weathered Coal Cliff Sandstone

lithology is slightly softer and erodes slowly. Hence the sandstone is cut back by a

number of processes resulting in the loosening and removing of small sheets of rock

surface and of individual crystal grains.

W h e n a unit has been deeply incised, the overlying units become unsupported and

ultimately blocks of the harder lithologies fall down. This undermining of the cliff

leads to a series of small collapses of individual blocks. These block collapses

c o m m e n c e at the base of the cliff and progressively m o v e upwards, or alternatively the

entire cliff face can collapse at one time as occurred in the Coalcliff area in 1988. T h e

physical processes of erosion which are operating on the cliffs are fairly easy to

delineate and define. At least five processes are thought to be acting on the cliffs and

causing differential erosion in this critical location. These processes are outlined below,

along with the effect that they appear to have on the rocks.

(1) W a v e action causes hydraulic pumping at sea level, giving alternately high and

low pressure in joints due to direct wave action. It is probably more effective

than abrasion, especially where closely jointed coal and claystone are present at

water level.

(2) Direct physical abrasion by seawater laden with sand and small pebbles. Running

water has an acknowledged effect on rocks of all kinds and this abrasion could

be a major feature in the cliff decay. It is likely that the more particles carried
108

by the sea, the greater would be erosive effect of the waves. Storm waves are

likely to do more erosion in a given period of time than do normal sea waves.

The Coal Cliff Sandstone is the most susceptible to this battering, particularly

when it occurs at sea-level (Fig. 5.25).

Related to waves, is sea spray tossed up when the waves break. This spray often

travels long distances and could, over extremely long periods of time, erode a soft

rock unit by salt crystallisation and thus cause it to recede. During storm periods

the spray is thrown faster and farther, and often it is accompanied by salt

weathering up to several hundred metres above sea level.

Another feature is the marine cyclic wetting and drying that occurs between

regular tides and between seasonal storms. Simple tests on soft rocks show that

wetting a sample after a cycle of drying and heating will greatly increase the rate

of breakdown. Wetting and drying is probably very effective and accounts for the

development of most of the shore platform in the area. Between successive high

tides the preserved portions of the platforms remain saturated whereas the rock

above mid-tide level is subjected to wetting and drying. This allows the

development of a notch at the base of the cliff which promotes cliff retreat by

collapse. Similar processes can act on the exposed escarpment due to rainfall.

The sea level at times during the Quaternary was at least several metres higher

than its present level. The lower units of the Narrabeen Group including the

Wombarra Shale in the north and much of the Illawarra Coal Measures would

have been either below sea level or within the zone of wave breaking. Active

erosion and weathering would have occurred in the low strength portions of these

units.

During rain storms, there is an increase in the volume of water getting into the

surface pores of the rocks. The Coal Cliff Sandstone is slightly porous but they

are permeable through secondary openings such as joints. As weathering advances


109

the porosity of a rock generally increases due to removal of cementing materials.

Into the resulting openings, water could be splashed and on drying salts would

crystallise out. With time and repeated cycles of wetting and evaporation, the salt

could build up and force grains off the sandstone surface. The Coal Cliff

Sandstone appears to be susceptible to such salt build up probably become it is

close to sea level through m u c h of the area. Also where the content of clays in

a rock is high, destruction caused by expansion and contraction related to the

cyclical wetting and drying m a y occur, especially in the upper part of cliffs.

(6) The last process considered here is differential erosion caused by groundwater.

Groundwater can affect the Coal Cliff Sandstone where the water is channelled

d o w n joint planes and forced out along a diastem. Openings with different

heights, lengths and widths appear to have resulted from groundwater flowing

d o w n joints and along the interface between the Coal Cliff Sandstone and the

underlying laminated sandy shale. Also in Bulgo and Scarborough water moves

laterally above clay aquicludes.

The petrology and mineralogy of the interbedded shale and claystone are such that the

expandable clay minerals present promote breakdown of the rocks. The most c o m m o n

cause is the influence of water, whether that influence be as a high pore pressure in a

joint complex in the weathered rocks (high artesian water pressures significantly

accelerating weathering, erosion and creep) or simply by saturating a rock mass and

increasing its weight. Solution of the siderite and calcite cements in the sandstones also

aids breakdown. The cyclic nature of the sedimentary sequence, specifically the

alternation of relatively permeable sandstone with impermeable claystone and shale

produces perched water tables and confined aquifers which invariably promote the

presence of significant pore water pressures. Artesian groundwater conditions, which

have been mentioned above, serve to illustrate the important influence of water in rock

failure.
110

On the basis of observations during this research, slides in talus often occur after rainfa

in the Illawarra area exceeds 350 mm per month and catastrophic slides invariably occur

when it is more than 450 mm.

5.9 SUMMARY AND CONCLUSION


Of all the geological factors influencing the stability of rock slopes, there is little doub

that the through-going faults and shear zones and the intersections of such structures are

the most significant. Because of their continuity they can influence large areas of a

rock slope. In addition, geologic displacement along faults and shear zones have led

to the crushing or overriding of most irregularities in at least one direction so that low

residual shear strength values are often applicable rather than the higher strength values

associated with more irregular rock surfaces.

Chemical alteration of the surrounding rock and the frequent presence of breccia and

clay gouge are also commonly associated with faults and shear zones. These factors

lead to a decrease in influence of surface irregularities as the intact material is more

readily sheared off. Finally, the presence of clay gouge adjacent to the polished or

smooth rock surfaces of faults can mean that the usually low strength encountered in

the laboratory for soil-rock surfaces is applicable to the field problem. In spite of their

size and continuity, the major faults and their intersections are not always readily seen

until after the slope failure develops.

The relationship of landslides to structural geology in the slip area appears to be very

important. These relationships may have direct and indirect effects on the landslide

problem. The direct effects occur in areas such as faults zones, where water movement

can give rise to major stability problems; the indirect effects are due to the structural

pattern causing certain beds to may form unstable foundations (for example the

Wombarra Shale occurs at road and railway levels). The best example of the direct

effects is the area of the Clifton Fault. At this point the road was completely cut in
Ill

1988 and remained so for a long period. The road from the north rises through a

cutting in the Scarborough Sandstone until the fault zone is reached. South of the fault

the m o v e m e n t has brought the basal portion of the Wombarra Shale against the fault at

road level. At rail level the northern and southern faces of the fault zone are occupied

by the Stanwell Park Claystone and Wombarra Shale respectively. The fault-zone also

acts as a drainage channel. Weathering and erosion along it have resulted in a thicker

talus cover than in nearby areas. The fault acts as a feeder for underground water and

even after prolonged dry spells water still runs from the area. The net result is that a

relatively small rainfall can thoroughly saturate the talus in the fault zone, where it is

already in a highly unstable position. This fault causes an increase in local water flow

and appears to be directly related to the Moronga Park slump in the area. The Jetty

Fault between Clifton and Coalcliff also appears to be responsible for the slide k n o w n

as the Jetty rock slump. The fault acts as a drainage channel for groundwater

circulation under the road. Similarly the Harbour Fault, south of Coalcliff, affects land

stability in the slip area and appears to be responsible for the slide k n o w n as the

Harbour slump.

The stabilities of the cliffs are mostly affected by the steeply dipping joints and by the

differential erosion of sandstone, shale and claystone. Bedding and cross-stratification

also affect the rate of retreat, and the shape of the blocks. The overall retreat rate is

quite slow, but there are occasional zones which suffer short-term, sudden movements.

The lithology forming the cliffs in the study area is the lower Narrabeen Group. This

can be divided in the field into six basic lithological types comprising Coal Cliff

Sandstone, W o m b a r r a Shale, Otford Sandstone M e m b e r , Scarborough Sandstone,

StanweU Park Claystone and Bulgo Sandstone. Sandstones are the units most resistant

to differential erosion and generally define the line of cliffs. The W o m b a r r a Shale and

Stanwell Park Claystone comprises the weak rocks (soft, fractured and weathered rocks)

which are subject to differential erosion. They cause cliff collapse by undercutting
112

(weathering and erosion). Interbedded layers of shale are commonly associated with the

Coal Cliff Sandstone, Scarborough Sandstone and Bulgo Sandstone and provide minor

zones of weakness.

Lithologic control of joint types and spacing, including their orientation, planeness and

the existence of certain joints, is important in the study area. Joints were studied in th

various sandstone, shale and claystone units. Joints systems, which are well developed

in sandstones, may also occur in claystones and shales but the latter units usually

contain a great or number of joints. Joints in sandstones pass downward into the

claystones and shales with little change in strike or surface characteristics but a sligh

increase in frequency. Joints are parallel and of similar planeness in shales and

sandstones but are more widely spaced in sandstones. A sharp discontinuity in joint

spacing occurs at the contact between shale, claystone and overlying sandstone. The

joints in claystone and shale are spaced at intervals of fractions of millimetres to seve

centimetres, whereas joints in the overlying sandstone are at intervals of several

centimetres to many metres.

Lithologic control of joint characteristics is apparent throughout the stratigraphic sect

Joints in the Scarborough Sandstone are typically widely spaced (1-4 m) whereas in the

interbedded sandstone and siltstone of Bulgo Sandstone they usually show a 0.5-1.5 m

spacing. Thus the widely spaced joints divide the Scarborough Sandstone into big

rectangular blocks while closely spaced joints divide the Bulgo Sandstone into moderate

to small blocks. As a result, more rockfalls occur from the Bulgo Sandstone than from

the Scarborough Sandstone in the study area. However, falls that occur in the

Scarborough Sandstone are usually much larger and more destructive.

The high horizontal stress environment known to exist in the Illawarra area is an

important factor which influences slope failure. In the geologic past, and to a certain

extent now, the high inherent confining stresses produced stress relief joints in the

vicinity of the escarpment. Stress relaxation would be at a maximum on the very edge
113

of the scarp with the ensuing fracturing contributing to the instability of the slope as

follows.

(1) Stress relief loosens and weakens rock in escarpment walls and thereby enhances

weathering and erosion.

(2) Escarpment development may proceed independently of tectonic processes via

progressive stress relief fracturing and erosion of escarpment bottom rock.

(3) Stress relief fracturing increases rock mass permeability and deformability and

decreases rock mass strength. Groundwater flow through stress relief fractures in

the escarpment walls contributes to weathering and colluvium development.

(4) Stress relief fracture permeability and variations in lithologic layering control

groundwater flow in escarpment walls and bottom. Therefore, artesian groundwater

systems may occur and contribute to escarpment bottom heave and fracturing.

Escarpment bottom heave and fracturing, along with groundwater flow, contribute

to talus instability.

(5) Stress relief tension joints and bedding planes are also involved in colluvium

development, as well as rock falls and rock-block creep on escarpment walls.

(6) Stress relief fracturing expedites rock weathering, alteration and solution and

influences groundwater flow as mentioned above. It influences the stability of

natural, excavated slopes, tunnels and also the stability and deformation of

foundations placed on rock in the destressed zone along the escarpment wall.

An analysis of the joints cutting the sandstone was performed to determine if they had

any preferred orientations. The resultant data show that major joint directions indicate

stress directions compatible with the probable stress directions present during the

formation of major structures in the Sydney Basin. The resulting joint strike maxima

for the lower Narrabeen Group show that the most prominent joint set exposed at the

surface, with a direction between 005 and 025, has a significant effect on slope

stability in the study area.


114
115

CHAPTER 6
REVIEW OF THE ROLE OF GROUNDWATER, RAINFALL,
HYDROGEOLOGY AND EARTHQUAKES

6.1 GROUNDWATER
In most large landslides groundwater has an important influence in controlling stability.

The most important influence is the reduction in shear strength along the slip surface

due to an increase in pore water pressure. Water flow also creates seepage pressures

which usually act in the direction of flow or seepage. Thus groundwater flow can play

a significant role in the development or triggering of landslides. Moreover, seepage

pressures may facilitate external and internal erosion in soils and weathered or closely

jointed rocks (Coates, 1990).

6.2 INTRINSIC PROPERTIES

Porosity and permeability are the two important factors governing the accumulation,

migration and distribution of water in soils and rocks. However, both may change

within a rock or soil mass in the course of its geological or weathering evolution.

Furthermore, it is not uncommon to find changes in both porosity and permeability with

depth due to variation in a number of features, including pore size distribution. The

actual size of the pores is significant since in narrow pores or capillaries surface tension

forces exert a control over the movement of fluids. In addition, chemical interaction

may occur between the water and dissolved gases and certain rock or soil constituents,

particularly clay and soluble minerals such as carbonate cement.


116

6.2.1 POROSITY
The factors affecting the porosity of a rock include particle size distribution, sorting,

grain shape, fabric, degree of compaction and cementation, solution effects and lastly

mineralogical composition, particularly the presence of clay particles.

The highest porosity is commonly attained when all the grains are the same size. The

addition of grains of different sizes to such an assemblage lowers its porosity and this

is, within certain limits, directly proportional to the amount added. Irregularities in

grain shape also result in a large possible range of porosity, as irregular grains may

theoretically be packed either more tightly or loosely than spheres. Similarly angular

grains may either cause an increase or decrease in porosity.

After a sediment has been buried and indurated, several additional factors help determine

its porosity. The chief amongst these are closer spacing of grains, deformation and

crushing of grains, recrystallisation, secondary growth of minerals, cementation and, in

some cases, dissolution. Hence the diagenetic changes undergone by a rock may either

increase or decrease its original porosity.

The porosity of a deposit does not necessarily provide an indication of the amount of

water that can be obtained therefrom. Even though a rock or soil may be saturated,

only a certain proportion of water can be removed by drainage under gravity or by

pumping. The remainder of the water is held in place by capillary or molecular forces.

6.2.2 PERMEABILITY

In ordinary hydraulic usage, a substance is termed permeable when it permits the

passage of a measurable quantity of fluid in a finite period of time, and impermeable

when the rate at which it transmits that fluid is slow enough to be negligible under

existing temperature-pressure conditions.

The flow through a unit cross-section of material is modified by temperature, hydraulic

gradient and the permeability. The latter is affected by the uniformity and range of
117

grain sizes, shape of the grains, size and shape of the pore throats stratification, the

amount of consolidation and cementation undergone and the presence and nature of

discontinuities.

The permeability of a rock or soil material is strongly affected by the interconnections

between the pore spaces. If these are highly tortuous then the permeability is

accordingly reduced. Consequently tortuosity figures are important in permeability

considerations, since they influence the extent and rate of free water movement.

Tortuosity can be defined as the ratio of the total path covered by a current flowing in

the pore channels between two given points to the straight line distance between them.

Stratification in a formation varies within limits both vertically and horizontally. It is

frequently difficult to predict what effect stratification has on the permeability of the

beds. Nevertheless in the great majority of cases where a directional difference in

permeability exists, the greater permeability is parallel to the bedding.

6.2.3 RELATIONSHIP BETWEEN POROSITY AND PERMEABILITY


Porosity and permeability are not necessarily as closely related as would be expected.

For instance, very fine textured sandstone frequently has a higher porosity than coarse

sandstone though the latter may be much more permeable. In other words, the size and

interconnectedness of the pores is all important as far as the permeability of a formation

is concerned. It is not uncommon to find variations in both porosity and permeability

throughout a formation.

6.2.4 FRACTURE (SECONDARY) PERMEABILITY


The permeability of intact rock sample does not necessarily bear any relation to the

permeability of the rock mass. The permeability of intact rock (primary permeability)

is usually several orders magnitude less than the in situ permeability (fracture

permeability). Although the secondary permeability is affected by the frequency


118

(spacing), continuity, openness, degree of interconnection of discontinuities and amount

of infilling of discontinuities, a rough estimate of the permeability can be obtained from

their frequency. Admittedly such estimates must be treated with caution and cannot be

applied to rocks which are susceptible to solution. In rock masses the discontinuities

are the most important conduits for water movement. Discontinuities allow water to

percolate through rocks with extremely low values of porosity. Indeed the frictional

resistance to flow through discontinuities is frequently much less than that offered by

a porous medium. Hence appreciable quantities of water may be transmitted.

6.3 POREWATER PRESSURE


Long-term landslides are not triggered by a rapid increase in shear stress, but either by

a slow, progressive decrease in shear strength parameters or commonly by an increase

in porewater pressure which also has the effect of reducing shearing resistance within

the slope (Crozier, 1986). Porewater pressure may be positive, resulting from a build-

up of groundwater above the shear plane, in which case the normal stresses are reduced.

Positive porewater pressure at any point in a freely draining slope is determined by the

product of the height of the water table or, more accurately, the piezometric surface,

vertically above that point and the unit weight of water. It exerts an upthrust which,

by reducing normal stress, reduces the resistance within the slope. This upthrust will

be enhanced if groundwater within the slope is under artesian pressure. Under

unsaturated conditions, porewater pressure may have a negative value, resulting from

tension exerted by attached water, and hence may provide an increment of strength.

If groundwater is already within the zone subject to applied loads or surcharge, sudden

loading will prevent drainage and excess porewater pressure will develop, immediately

reducing resistance. Hutchinson and Bhandari (1971) have described such 'undrained

loading' resulting from debris accumulating on the head of a mudslide as an important

mechanism in promoting downslope movement. Their work was prompted by the


119

observation that certain coastal mudslides advanced on slopes flatter than those

corresponding to the limiting equilibrium for conditions of residual strength and

groundwater flowing parallel to the slope surface. From both theoretical considerations

and porewater pressure measurement in the field, they determined that debris

accumulating at the rear of a mudslide loaded the landslide sufficiently quickly to

prevent drainage and to increase porewater pressure to artesian levels. Such pressure

appears to induce a forward thrust which may initiate shearing movement or accelerate

the rate of movement downslope.

6.4 CHANGES IN WATER CONTENT


Changes in water content can quickly affect the stability of slope materials and have

been responsible for triggering, reinitiating and accelerating more landslide, than any

other factor. In nearly all cases, an increase in water content decreases stability in one

or more of the following ways.

(1) Increasing interstitial porewater pressure. Positive porewater pressure, which

reduces resistance, can be developed within the phreatic zone or within

groundwater zones perched above a relatively impermeable substrate.

(2) Developing cleft-water pressure within joints, voids and fissures. This has a

similar effect on resistance to that produced by interstitial porewater pressure.

(3) Developing seepage pressure where a drag stress is set up in the direction of water

percolation, thus contributing to shear stress. Seepage pressure may also lead to

gradual subterranean erosion and removal of underlying support.

(4) Increasing weight. The effect of an increase in weight provided by water is to

increase the disturbing forces along a potential slip surface. Increase of weight

can only instigate failure where the slope is already close to the critical

equilibrium.
120

Slopes consisting of clay, particularly the expanding-lattice variety, are the most

susceptible to failure though the loss of cohesion. Some cohesion soils can take

up large amounts of water, and commonly exhibit desiccation cracking which

facilitates the entry of water.

(5) Decreasing cohesion (apparent cohesion). In soils, cohesion results mainly from

capillary and electro-molecular forces which are reduced as the amount of

interstitial water increases. In weathered granite saprolite, for example, Lumb

(1966) noted that apparent cohesion can be as high as 200 kPa but that it will

drop to zero when the material is saturated.

6.5 EFFECTS OF SOLUTION


The movement of groundwater within rock masses containing a significant proportion

of soluble minerals exerts a major control over their removal in solution. The actual

site of dissolution is often controlled by the presence of rock mass discontinuities and

other variations in mass permeability, including low permeability clay seams. However,

besides removal of soluble materials from surfaces, reductions in bulk density or cavity

formation may also be prevalent. The amount of mineral removed from the rock mass

depends mainly on the rate of flow, quantity, temperature and chemistry of the water

passing through it. Also of major significance is the solubility characteristics of the

rock forming mineral.

6.6 GROUNDWATER FLOW IN SLOPE STABILITY PROBLEMS


In slope stability problems only a portion of the regional flow system remains of interes

- that portion of the flow system which occurs within and adjacent to the slope. In

previous geotechnical literature the most common way to portray the groundwater flow

was to show the flow occurring subparallel to the groundwater table which could be

delineated by the groundwater level encountered by borings penetrating the slope (Fig.
121

6.1). Such a portrayal of groundwater flow within a slope is quite different from the

present view of the general case for groundwater flow in slopes (Fig. 6.2).

One significant result of the recent developments in the theory of groundwater flow

systems applicable to slope stability is that there is normally a downward pore-pressure

gradient in holes drilled in the upper portion of a slope and an upward pore-pressure

gradient in holes drilled in the lower portions of slopes (Fig. 6.1b); assuming

homogeneous and isotropic permeability.

Significant variations from this would occur in areas where there is a regional

groundwater recharge or discharge, and where the permeability within, or recharge to,

the slope is non-uniform. Perhaps the greatest difference between the two portrayals of

groundwater flows in Fig. 6.1 occurs in the discharge area; an area of considerable

interest in the slope stability problem. According to the first case shown in Fig. 6.1a,

no adverse groundwater flow conditions are likely to result from the placing of an

impervious fill at the base of the slope. This is because the flow is parallel to the

surface and, therefore, the fill has no appreciable influence on the flow of groundwater

in the region of the base of slope. However, it is obvious from the second case (Fig.

6.1b) that placing an impervious fill at the toe of the slope would result in an

appreciable disruption of the groundwater flow within the hill and a build-up of

porewater pressure.

6.7 SLOPES COVERED WITH LANDSLIDE DEBRIS


Most geologists have observed that landslide debris is frequently wet and unstable,

usually much less stable than the original slope and adjacent slopes. Fig. 6.2 illustrates

the manner in which the slide debris can block the normal groundwater discharge area

of the slope to produce a characteristically unstable deposit of slide debris. The

equipotential lines and flow lines shown in Fig. 6.2a illustrate the distribution of

groundwater pressures and the flow of water within a slope before a slide develops.
122

Fig. 6.2b indicates the case where the slide debris covering the slope increases the level

of the groundwater table and increases the groundwater pressures in the area of the slide

debris.

A similar effect can occur in slopes formed in thinly bedded materials where adjacent

layers possess quite variable permeabilities. A shearing displacement can cut off the

outlet for the groundwater flow within the hill and allow porewater pressure to build up

to critical values more readily following the initial displacement than before the slide

began. The overall effect of either phenomena is to accelerate the m o v e m e n t of the

slide and will tend to remove the slide debris from the hillside in a relatively short

period of geologic time. This type of groundwater behaviour is also one of the main

factors which cause deposits of slide debris to be so unstable, even though they have

m u c h flatter slopes than the original slope.

6.8 HIGH WATER PRESSURES IN THE ESCARPMENT

The significance of high water pressures beneath escarpments and in escarpment walls

has not been widely recognised for several reasons:

(1) very few piezometer installations have been made which are extensive and deep

enough to illustrate the phenomena;

(2) significant features can be masked by the effects caused by landslide debris;

(3) the groundwater discharges (springs) tend to occur at the base of the escarpment

where they are not noticed; and

(4) in areas of harder rocks not prone to landslides, and where better exposures are

available, the influence and effects of the regional groundwater discharge are

minimal.

Certain geologic environments (flat lying and inclined layered rocks with great difference

in their permeability) provide conditions for high porewater pressures to develop at the

base of an escarpment. In particular, the presence of thick low permeability rocks, such
123

as shales and related clay-rich rocks, volcanic ash deposits, thick fault zones and buried

soil profiles, would tend to be associated with zones of excess porewater pressure within

and at the base of escarpment slopes.

A schematic diagram showing the various effects that can develop due to the presence

of groundwater discharge at the base of an escarpment is given in Fig. 6.3. The high

porewater pressures can also act on pre-existing joints and bedding planes to decrease

the stability of the slopes and could lead to widespread landslides.

6.9 SPECIAL EFFECTS OF FAULTS ON THE HYDROGEOLOGY OF SLOPES


A significant engineering property of faults is their effect upon the permeability of a

rock mass. The typical composite fault that was shown in (Fig. 5.1) m a y have one or

more low permeability zones associated with the fault gouge which separates two zones

of high permeability in the fractured rock (e.g. Fig. 6.4a). In addition, the fault breccia

(if present) m a y be more permeable than the gouge (Fig. 6.4b, c). Thus, faults can act

as groundwater barriers as well as groundwater conduits, or as both at the same time.

The net result of this complex permeability layering within a fault zone is that faults

can have a variety of effect on the flow of groundwater and the resulting distribution

of fluid pressures has a major effects on the stability of a slope. It is not u n c o m m o n

to find springs and seepage of groundwater along faults which have served as a failure

surface in open pit or underground mines. Wilson (1959) described such conditions

for a failure of a portion of the Bingham Canyon pit.

Several consequences of this zonation are illustrated in (Fig. 6.4). O n e possibility is

that the fault m a y act as a groundwater barrier as shown in (Fig. 6.4a). In this case the

rock adjacent to the pit slope m a y be well drained yet unfavourable groundwater

conditions m a y exist that could lead to a slope failure. Fig. 6.4b shows a fault serving

as a groundwater conduit leading water from a nearby stream into the pit slope. In this

case the dual behaviour of the fault due to the presence of one or more low
124

permeability layers in addition to the fractured rock may prevent the groundwater in the

fault zone from reaching the drainage gallery shown. Fig. 6.4c shows a fault serving

as a subsurface drain which would increase the stability of the mine slope. These

conditions are equally applicable for groundwater condition adjacent to a natural

escarpment or artificial cutting.

6.10 HYDROGEOLOGICAL ASPECTS OF THE ESCARPMENT IN THE STUDY

AREA
In a consideration of the hydrogeological aspects of the northern niawarra slip area the

conditions in each of the Hawkesbury Sandstone, Narrabeen Group and Illawarra Coal

Measures are important. On the basis of field observations the Hawkesbury Sandstone

is quite porous, although the fracture porosity may be very localised. Along the

Lawrence Hargrave Drive between Clifton and Coalcliff, the Scarborough Sandstone and

Bulgo Sandstone also show high fracture porosity along the joints (Fig. 6.5).

Along the outcrop of the Illawarra Coal Measures the coal seams have been observed

by the author to be permeable and to act as aquifers. The presence of aquifers, unstable

cements and swelling clays contribute to rapid weathering and disintegration and reduce

the overall rock mass quality in the Illawarra Coal Measures in outcrop.

Rainwater enters the various Narrabeen Group aquifers where they intersect a surface

recharge zone, be that along creek beds on the plateau or on the slopes of the

escarpment and under scree slopes. The plateau areas act as the major catchment to

replenish the groundwater table within the escarpment slope. However, sandstone

aquifers along the edge of the escarpment would be exposed to recharge from the

escarpment slopes. The area west of Coledale Station on the Illawarra Railway is an

example of a mid-escarpment recharge zone (Fig. 6.6a, b). A series of perched and

confined aquifers are also present in the escarpment because most of the claystone

sequences within the Narrabeen Group are relatively impermeable. The primary porosity
125

and permeability of the Narrabeen Group is very low. It is considered that the vast

majority of the groundwater is stored in fissures and fracture systems such as faults,

dykes, joints and bedding plane partings. Lateral facies changes would also inhibit the

horizontal flow of groundwater via cracks to some extent. However, water is

transmitted via these secondary discontinuities usually as a vertical flow over limited

distances until a relatively thick impermeable claystone bed was encountered. In the

cliffs between Coalcliff and Clifton water was observed by the author to be issuing at

an estimated rate of about 8 litres per hour from the base of the Scarborough Sandstone

(Fig. 6.5). The outlet pipes were roughly spaced every 1 m and this was several days

after moderately heavy rainfall. This confirms the impermeable nature of the underlying

Wombarra Shale at this locality. Water is commonly seen issuing at other locations

along the escarpment. For example, between Stanwell Park and Coalcliff, where water

is flowing from the weathered Wombarra Shale and Coal Cliff Sandstone, or between

Scarborough and Clifton where water is flowing from the colluvium and over buried

Wombarra Shale (Fig. 6.10). It would be expected that the rocks in the immediate

vicinity of the escarpment would have a higher permeability due to stress relief joints

and weathering of the joint and bedding plane partings.

The relatively common observation of water issuing from the coastal cliffs and

escarpment slopes is most likely explained by discontinuities in the rock mass. The

sandstone of the Narrabeen Group appears quite permeable although the aquifers are

sporadically and irregularly spaced. Stress relief with the opening of discontinuities

would be most significant on the edge of the escarpment. The Southern Coalfield is

intersected by faults, with throws of up to 100 m, and by intrusive dykes. None of

these structures yield any significant amount of water into the mine workings. This is

most likely due to the alteration and the presence of impermeable clay lined fault zones

and dykes and to the relatively impermeable nature of the claystone and shale units,

especially the Wombarra Shale.


126

Many of the slopes are blanketed with relatively impervious clayey talus which covers

horizontally bedded strata in the study area. High groundwater pore pressures and

artesian conditions develop and cause failure of the talus, whose shear strength

approximates its residual value. For example, a deep seated failure has been active for

some time at Coledale and was moving slowly toward the sea in this place (Fig. 6.6c).

This was exacerbated by the heavy rainfall in April 1988 (Figs 6.7 to 6.9). This failure

was most probably controlled by a weak colluvial soil layer beneath the fill embankment

and overlying the bedrock (Mostyn and Alder, 1991). The Coalcliff slump between

Coalcliff and Stanwell Park is another example; this slump is active and will be

explained in chapter 8.

The effect of Water Board dams, which are located to the west of the escarpment, is

important for the slope stability in some places. For example, Coalcliff dam is 1.7 km

west of the railway at Coalcliff and at an elevation of 210 m above the railway level

(Fig. 1.4). The gully along which the dam is situated appears to correlate with the large

north-trending fault zone which includes the Ladysmith and Western Gully Faults. In

view of the probability that subsidence associated with the extensive coal mining in this

area would have caused disturbance to the rock, sufficient to cause opening up of pre-

existing fault related fractures, it is probable that the rock beneath the catchment and

storage area would have a greater than normal vertical permeability. Figure 6.11 (also

see Fig. 6.4c) illustrates this hypothesis.

It is thought that this type of groundwater infiltration has probably increased due to coa

mine subsidence between 1967 and 1973. This would have caused a significant increase

in the water flow in the vicinity of the escarpment and also raised the regional

groundwater level.
127

6.11 RAINFALL A N D ITS RELATIONSHIP T O H Y D R O G E O L O G Y

Field studies and previous investigations suggest that groundwater flows, with or without

the contribution of local surface infiltration, may be a major cause of the slope

instability in the northern Illawarra slip area. Consequently, this factor has received

considerable investigation with several dozen piezometers having been installed

throughout the area of slip over the last ten years. Evidence of artesian groundwater

flows was discovered in the region (Coalcliff area, at between 5 and 6.5 m depth; State

Railway Authority, 1982). It would appear that the water is flowing within the top part

of the Stanwell Park Claystone. The layer was isolated above and below by grey shale

beds.

Correlation of rainfall with groundwater levels and slip movement has been attempted

during this and previous investigations, however no obvious correlation was found

linking short term events. During the flood rains of April 1988 along Lawrence

Hargrave Drive between Coalcliff and Clifton, large quantities of talus and topsoil from

above the road slipped down toward the ocean at six locations. Rainfall records at

Wombarra and Stanwell Park showed falls of about 656 mm for the 3 days of April,

just before the slip, which came on top of the earlier rainfall in the same month that

had saturated the area.

Young (1978), in attempting to predict rainfalls likely to initiate landslip in the Illawarra

area, stated that in Wollongong the recurrence interval of the most intense storms bears

little relation to the frequency of severe landslip activity. This proposition appears to

be supported by surveys of the rainfall and historical records which show that the

periods of greatest slope instability in the Illawarra area coincide with the years with

the highest cumulative rainfall, not necessarily the years with the most intense storms.

For example, Young (1978) noted that, for the Illawarra area in general, daily rainfall

exceeding 400 mm/day has occurred as follows :


128

10/2/1975 493 mm

11/9/1950 433 mm

15/3/1936 508 mm

05/5/1925 510 mm

13/1/1911 443 to 529 mm

The point to note is that although 1950 did not experience the most intensive storms,

it did have the highest yearly rainfall on record, and this year was noted for its larg

number of slope failures and, in particular, major slipping at Coalcliff. It is also

interesting to note that, according to Young (1978), the storm on the 11/9/1950, despite

being the most intense of the year, did not appear to cause any specific slope

movements. For the Illawarra area during this study annual rainfall has been low and

daily rainfall exceeding 100 mm/day has only occurred as follows:

07/6/1991 110 mm

11/6/1991 225 mm

12/6/1991 168 mm

09/2/1992 101 mm

10/2/1992 240 mm

22/3/1992 104 mm

14/9/1993 102 mm

6.12 RAINFALL AND ITS RELATIONSHIP TO LAND MOVEMENTS


The topography of the Illawarra escarpment tends to encourage rainfall. It rises steepl

from the sea some 300 m to the upper plateau over a small horizontal distance. As a

result heavily laden clouds travelling inland from the sea lose rain as they pass over

scarp. Geomorphological features highlight periods of intense past surface water flows.

Erosion of the main gullies, particularly noticeable above the main cascades at the top

of the escarpment, has occurred in at least two cycles. Adamson (1974) described how
129

the western and eastern gullies above Coalcliff had excavated the main bulk of the

valleys in the first cycle of erosion after which a lower flow situation occurred which

created a narrow entrenched stream in the base of the gully.

Hydrograph data shows that rainfall is highly variable along the escarpment but is

commonly concentrated within isolated areas on the steeper slopes. This leads to high

surface water flow and infiltration. Annual average rainfall along the coastal plain

ranges from 1100 - 1200 mm but exceeds 1700 mm (Fig. 2.4) at the crest of the

escarpment. Records are available for 1950 showing more than 3000 mm along part

of the escarpment at Mt Keira (Young, 1978).

In contrast to Young (1978), Brand (1993) and Olivier et al. (1994) documented that

mass slope failures appear to have occurred during high intensity rainfall events of more

than 400 mm over 24 hours, which occurred within a longer duration rainfall period.

Individual high intensity rainfall events such as in June 1991 and February 1992 (Figs

6.7 and 6.8) appear to have caused scours and flooding rather than activating major

landslips.

Bowman (1972) found no correlation of landslip with daily rainfall events, however he

postulated a connection of major failure with monthly rainfall in excess of 400 mm.

This was confirmed by Young (1978) and she suggested a monthly rainfall in excess

of 250 mm as likely to trigger landslip. She also noted a lag time between heavy

rainfall and accelerated movement. By studying rainfall and evaporation data, she

compared excess precipitation levels for the low coastal plain (Albion Park) with the

upper levels of the escarpment (Mt Keira). She concluded that on the higher slope the

average quantity of water available for runoff and infiltration is 55% of the average

rainfall, as compared to 15% in the lower slopes. This is even more dramatic when it

is considered that the upper levels receive more rainfall than the lower regions.

In reality, any prediction of movement with respect to rainfall is dependent upon

infiltration levels and the reaction of phreatic surfaces to cumulative rainfall events. As
130

the phreatic surface rises to a threshold level, movement occurs. This level represents

a situation where the factor of safety is unity. Longmac Associates (1991, see also Fig.

6.12 and Table 6.1) studied the relationship of monthly rainfall with respect to landslid

activity and found a poor correlation existed for the 1988 - 1992 period. They

concluded that data for three monthly rainfall records (e.g. 1950 with 1171 mm per 3

months) correlated better to the current large scale landslide (Moronga Park) in the

detailed study area. Fig. 6.12 depicts rainfall versus recurrence interval highlighting t

major rainfall events.

Between 1988 and 1990 was an active period for landslip occurrences. Referring to the

rainfall data (Table 6.1 and Figs 6.7 and 6.8), 1988 involved both high intensity and

long duration rainfall events. On the other hand, 1989 and 1990 displayed low monthly

and high three monthly levels. Landslip was equally active in each year. The year

1991 was mainly dry with only one major rainfall event in June (Figs 6.7 to 6.9; Fig.

6.9 shows a positive correlation between rainfall in the Clifton and Wollongong areas).

This resulted in many new debris flows but did not activate the existing larger landslips

The writer installed in late 1991 a series of surface survey pegs on the four landslides

(Clifton Hotel earth-slump, Moronga Park slump-earth flow, Jetty slump and Harbour

slump) for measuring slip movement, but no movement occurred during the study

period because of very dry conditions.

6.12.1 THE CONCEPT OF THRESHOLDS


Because of the relationship between the occurrence of landslides and rainfall events,

there have been numerous attempts to derive thresholds beyond which slopes will

become unstable, the objective being to establish a value of rainfall which, when

exceeded, results in landslides (Guidicini et al, 1977; Crozier, 1986; Bhandari et al,

1991; Olivier et al, 1994). Guidicini et al. (1977) analysed rainfall records for nine

regions in Brazil and the results indicated that when rainfall exceeded 12% of the mean
131

annual rainfall, a critical level of soil saturation was reached which in most cases

triggered landslides. When the intensity of rainfall was greater than 20% it appeared

that catastrophic landslides occurred.

General observations by the author over the 1988 - 1994 period have confirmed previous

deductions that land instability occurs after prolonged rainfall which acts to top up the

phreatic surface until a critical threshold is reached (more than 400 mm per month,

when rainfall exceeded 25% of the mean annual rainfall). This contrasts with an

isolated high intensity event during low rainfall periods which generally tends to cause

flooding and failure by scour or debris flow.

6.13 GROUNDWATER AND ITS RELATIONSHIP TO LAND MOVEMENTS


All of the rainfall evidence, when considered as a whole, tends to suggest that only

longer term rainfall trends on a scale of several months to years, appear to significantly

influence the output of aquifers affluxing beneath the slip and hence the groundwater

levels in the adjacent talus and embankment materials. Within the Narrabeen Group,

groundwater flow from weathered porous sandstone aquifers into the interface between

the bedrock and overlying talus would be expected to apply hydrostatic pressures of at

least 60 kPa to the talus at a depth of 6 m. The important factor to consider in this

case is that the groundwater pressure applied to the lower boundary of the talus resulted

from upward flow from beneath and to the west, and it is not simply due to the height

of the column of water represented by the water table in the talus.

The possibility of the occurrence of serious slip movement during relatively dry periods

can be explained by the time lag between the previous major rainfaU period, capable

of significant aquifer recharge, and the emergence of the groundwater flows and pressure

at the aquifer/talus interface.


132
6.14 SUMMARY AND CONCLUSIONS
It is clear that groundwater flow in and around faults and similar features, such as dykes

and sills, requires special attention in slope stability studies. The physical properties o

the fault zone materials must be considered as well as any change in permeability or

change in physical properties due to an offsetting of lithologic units or other geologic

structures. Intersections of faults require additional attention as the jointing intensity

be much higher here and the effects of weathering much deeper.

Perched water tables have been found to be quite common in the study area because

the many claystone sequences within the Narrabeen Group are relatively impermeable.

This involves the collection of surface water in a basin of higher permeability material

above the regional water table. Infiltration of the water then results, forming a local

rise in the phreatic surface aggravating creep type slides. Depending upon the

catchment, large water volumes may be involved which can often lead to rapid mud

flow failures. The Coledale Rawson Street disaster (1988) is a notable example of this

failure mechanism.

The natural pore spaces in the soil or rock constitute the primary porosity which may

be filled with water. In cohesive soils this water can be further subgrouped into free

or combined water. In rock, the permeability is controlled by the interconnection of

pores and fractures. Unfractured claystone often has a low permeability due to its fine

grain size whereas sandstone with little intergranular cement would be expected to

possess a higher permeability.

Adamson (1974) discussed the relative permeability of the strata in the detailed study

area. The Hawkesbury Sandstone is the most permeable of the rocks, with Scarborough

Sandstone being reported to be the most permeable rock unit in the Narrabeen Group.

Water may also be contained within joints, fault zones or fractures which constitute the

fracture (secondary) porosity. This form of water results in seepage concentrations

which are often under a considerable head of pressure.


133

Theoretically, in strata which are sub-horizontally bedded and comprised of interbedded

sandstone and claystone, horizontal permeability is often m u c h greater than the vertical

permeability. The more impermeable claystone beds act as barriers to the flow. This,

however is susceptible to local influences, especially jointing.

In the detailed study area a complex hydrological environment exists. Mine plans

indicate that the area contains scattered faults, dykes and sills whereas weathered open-

jointed strata near the scarp face result from stress relief. Although m a n y fault and

dyke locations are unknown at the surface, they are believed to effect the water regimes

and create water concentrations, and thus high hydrostatic pressure, if confined by a

more impermeable outer layer. M a n y of these joints are noticeable on the surface in

the Bulgo Sandstone outcrop and they m a y be present, to a lesser extent, in the

underlying Stanwell Park Claystone unit. Several explanations exist for the weak, highly

jointed nature of the Bulgo Sandstone unit near the outer face of the escarpment. It is

likely that stress relief, mine subsidence and the intrinsic composition of the material

contributed to this characteristic. All these factors combine to provide high vertical

permeability often resulting in seepage concentrations.

With a relatively impermeable colluvium mantle existing at the base of the steeper

sandstone slope where it joints the claystone terrace, seepage is restricted in its passage

from the more permeable bedrock resulting in large hydrostatic water pressures. The

Bulgo Sandstone overlying the Stanwell Park Claystone is highly jointed. Water tends

to enter the rock strata along the top of the escarpment and upper slopes through open

joints, flowing downwards until a relatively impermeable layer is encountered. Water

also enter the Scarborough and Coal Cliff Sandstone though open joint system along the

escarpment and passes d o w n to the underlying shale units.

Water then travels horizontally and attempts to outlet on the outer face of the

escarpment at the base of the sandstone units. ff a deep colluvium deposit is situated
134

adjacent to the aquifer, the hydrostatic pressures will build up and m a y result in creep

movement or even a more active slope failure.

The Stanwell Park Claystone has been observed by the author to the north of Coalcliff

Station to possess high fracture porosity in surface outcrops. Flow rates of up to 24

litres/minute have been reported from the fractured unit at Coalcliff site (Ghobadi &

Pitsis, 1993). Borehole data for this unit show that, behind the escarpment, the rock

is m u c h less fractured (chapter 3.8.5).

Based on observations by the author and the previous studies, it appears that fracture

permeability is the most important feature of groundwater movement, with most of the

fractures occurring in areas of stress relief.

Variation in water content of the talus has a marked effect on its shearing resistance and

hence on its stability. The presence of some water m a y increase the appearance

cohesion in partially saturated soil, as against perfectly dry material. However, in

saturated soil pore water pressure due to groundwater and seepage will be positive and

shear strength will decrease with any increase in pore water pressure. T h e effect of

increasing the water content has been discussed by Terzaghi (1950) and m a n y others.

It acts in several ways, including:

(1) decreasing cohesion due to filling voids with water and expelling the air;

(2) increasing the weight and hence shearing stresses;

(3) possibly dissolving cementing materials; and

(4) causing a rise in the piezometric surface, involving increase of pore-water pressure

and decrease of shearing resistance.

It is quite obvious from studying the rainfall figures and periods of prevalence of

landslides that the most unstable periods are those when the rainfall is the highest over

an extended time interval.


135

On the basis of observations during the site visits, rockfalls are triggered by prolonged

rainfall and their behaviour is very similar to that of talus slides, which appear to occur

after cumulative rainfall exceeds some threshold (more than 400 mm per month).

Debris flows may be triggered by intense storms whereas most slope movement and

slides occur as a result of prolonged rainfall of certain magnitude (more 400 mm per

24 hours). Commonly a distinct time delay occurs between heavy rainfall and failure.

This time delay varies greatly, from merely a day to several months.

In any part of the escarpment the water may fall directly on it during rainfall, be

brought to it by surface drainage or be derived from underground sources. It is where

these source are combined that the maximum effects are felt, a good example being

along the zone of the Clifton Fault, which has been described previously is an active

aquifer system.

An important factor to consider in this area is that the groundwater pressure applied to

the lower boundary of the talus results from upward flow from beneath and is not

simply due to the height of the column of water represented by the water table in the

talus. Consequently, lowering of the water table in the talus cannot be expected to

produce the normally required decrease in hydrostatic and uplift pressures on the

talus/bedrock interface where the bedrock aquifer affluxes. This aspect is particularly

relevant to the failure mechanism and the choice of remedial measures.

6.15 EARTHQUAKES
Earthquakes reduce stability by imparting both a shearing stress and a reduction in

resistance to slope material. Earthquake wave propagation is thought to have three

principal effects on slip materials:

(1) the direct mechanical effect of horizontal acceleration which, at high shaking

intensity, may exceed acceleration due to gravity. This provides a temporary

increment to shearing stress which is sufficient on occasions to trigger landslides;


136

(2) cyclic loading in clays, sands and silts with weak inter-particle bonding. In

saturated material, seismic loading shifts the weight of particles from its granular

support onto the porewater, thereby increasing interstitial pressure, buoying up the

mass and causing liquefaction. Loose sands and sensitive clays are particularly

susceptible to liquefaction by earthquake shaking and other vibrations; and

(3) reduction of intergranular bonding afforded by cohesion and internal friction, due

to sudden shock, irrespective of the degree of saturation. This may lower the

shear strength of material towards its residual value. The effect is similar to that

experienced by a brick building when shaking breaks the bonds between mortar

and bricks. Although this response may not immediately initiate movement, it

serves to make the slope susceptible to future triggering activity.

6.15.1 EARTHQUAKES IN THE STUDY AREA


Compared to countries which are situated in active tectonic zones, such as Japan,

Turkey, Iran and Chile, Australia is considered to have only a small earthquake hazard.

Two significant earthquakes have taken place in the vicinity of study area since

recording began in 1909; they are the Robertson earthquake of 21 May 1961 and the

Picton earthquake of 9 March 1973, both of magnitude 5.5 on the Richter scale and

they both occurred at depths of about 20 km. Also a significant earthquake took place

at a more distance location, Newcastle. The Newcastle earthquake of 28 December

1989 had a magnitude 5.6 on the Richter scale (Melchers, 1990). This earthquake

occurred at depth of about 13 km. Recently an earthquake took place at Cessnock

(49 km from Newcastle), it occurred on 6 August 1994 with a magnitude of 5.3 on the

Richter scale. Earthquakes of this size occur on average about once every eighteen

months in Australia (McCue et al, 1990).


137
6.15.2 SECONDARY EFFECTS OF EARTHQUAKES
The effects of the Picton earthquake have been described by Denham (1976, 1979), and

those of the Robertson earthquake by Cleary and Doyle (1962). The isoseismal maps

are given in Figs 6.13 and 6.14. In both earthquakes the damage was confined to old

buildings (some more than 100 years old). For the Picton shock, which was felt over

an area of about 6000 km2, light damage was experience over about 4000 km2. This

consisted of damage to plaster, brickwork and the tops of chimneys where the heat from

fires had destroyed the adhesive properties of the mortar. The Newcastle earthquake

resulted in 12 deaths, hundreds of injuries and serious damage to, or destruction of,

thousands of homes and buildings. The isoseismal map is given in (Fig. 6.15). This

was the first time in written history that lives have been lost as the result of an

earthquake in Australia.

No reports were received of any sand boils during the Newcastle earthquake. The only

documented coseismic subsidence was on the southern abutment of the Stockton Bridge,

10 km north of Newcastle. The lack of surface faulting is not surprising for an

earthquake of this size and focal depth. In Australia during the last twenty years, five

faults have ruptured the surface following large, very shallow earthquakes in Western

Australia, South Australia and the Northern Territory. In eastern Australia there are at

least two recent earthquake scarps, in Victoria and Tasmania, but each predate written

history.

6.15.3 INTERPRETATION AND EFFECTS OF EARTHQUAKES AND


STRESS ENVIRONMENT
The Robertson, Picton and Newcastle earthquakes occurred at shallow depths as have

virtually all the other earthquakes reported in the Sydney Basin. It is well known that

shallow earthquakes are much more damaging than deeper earthquakes.


138

Any earthquakes occurring to the west of the Illawarra escarpment and at shallow depth

would have a very significant effect on the escarpment. High accelerations could well

be expected on the edge of the escarpment and on the edges of local benches on the

slopes of the scarp. Typical stress concentrations at the base of the escarpment are

likely, with possible tensile regions near the edge of the cliff tops. These very small

compressive forces and possible tensile zones are likely to result in stress relief whic

induces fracturing of the rock parallel to the escarpment. This phenomenon is common

throughout the world and has been noted by Mencel (1974). Thus the horizontal stress

perpendicular to the escarpment at the edge of the escarpment equals zero. Joints which

occur perpendicular to the escarpment are always closed, indicating the presence of a

horizontal stress parallel to the escarpment. The value of this horizontal stress is

difficult to estimate and would obviously increase with depth. It is likely that the join

which occur at mean orientations of 005 and 055 and produce a saw tooth plan on

the edge of the escarpment cause localised stress changes in the vicinity of the joints.

This could have a significant effect on stability as a wedging system may result.

Slow structural deformation would continue in the study area in the month following

the earthquake. The effect of earthquakes on slope stability are most likely to be

important in the study area when the earthquake occurs during a wet period which has

produced high porewater pressures and decreased shearing resistance.

Two additional aspects of the effect of earthquakes on mining may be considered.

Firstly, it is possible that subsidence due to mining may increase as a result of an

earthquake. A second aspect relates to damage to the mines themselves; water leakage

problems may be enhanced or the upper level faults or joints may result in a rupture

surface farther down. These factors will also affect the regional groundwater conditions

and may have implications for land stability in the escarpment area.
139

CHAPTER 7

ENGINEERING GEOLOGY

7.1 ENGINEERING PROPERTIES OF ROCKS IN THE LOWER NARRABEEN

GROUP
The factors which influence the engineering properties of rocks can be divided into

internal and external categories. The internal factors include the inherent properties of

rock itself, whilst the external factors are those of its environment at a particular point

in time.

As far as the internal factors are concerned the mineralogical composition and texture

are obviously important but planes of weakness within a rock and the degree of

weathering are frequently more important.

Over the last decade the influence of weathering on the engineering properties of many

igneous and sedimentary rocks, under dry and saturated conditions, has been investigated

by numerous authors (e.g. Gunsallus and Kulhawy, 1984; Dobereiner, 1986; Kembla

Coal & Coke Pty Ltd, 1990, 1991; Jeffrey and Shakoor, 1990; Olivier, 1990; McNally,

1993; Ghayoumian et al, 1993; Haney and Shakoor, 1994; Ghafoori et al, 1994). A

great deal of test data has been accumulated on engineering properties of fresh and

weathered rocks. It is very well known that weathering generally affects the structure

and behaviour of rock. As the degree of weathering increases, rocks become more

porous and weaker. Additionally, it is also very well documented that the presence of

water decreases the strength of rocks. However, in nature rocks are found under

varying degrees of saturation. Therefore, it is important to know not only the

weathering state of the rock mass but also its strength and durability in order to predict

the stability of a rock (Bell, 1993).


140
7.2 WEATHERING
Many definitions of the term weathering have appeared in the abundant literature which

covers this large and somewhat diverse subject (e.g. Merril 1897; Rieche, 1950; Keller,

1957; Dearman, 1974; Oilier, 1984; Turk and Dearman, 1986). Weathering includes the

processes of alteration of a rock occurring under the direct influence of the hydrospher

and the atmosphere at or near to the earth's surface. Weathering as a natural process

may be either mechanical or chemical or a combination of the two. Mechanical

weathering is dominated by physical breakdown with opening of discontinuities, the

development of new discontinuities and the opening of mineral grain boundaries and

cleavages. Decomposition is the dominant process in chemical weathering and leads

through stages of discolouration of the original fresh rock to decomposition of silicate

minerals to from clay minerals, with attendant opening of grain boundaries. Thus

chemical weathering leads to changes in colour, strength and porosity, but original rock

mineral texture is largely retained until the final stages when a residual soil is forme

7.2.1 ENVIRONMENTAL FACTORS CONTROLLING ROCK WEATHERING


The influence of weathering on the engineering properties of rocks is well known,

because of its importance. The natural weathering processes described are closely

related to three important environmental factors; the hydrosphere, the climate and the

topographical situation (Fookes et al, 1988). The influence which the hydrosphere has

on the weathering processes is evident from the vital role that water plays in the

degradation processes outlined above. The presence of water in the rocks also affects

their mechanical properties. In nature the degree of saturation will depend on the

position of the rock mass relative to the groundwater table. The position of the

watertable, and moisture content in the overlying vadose zone, both vary seasonally.

Consequently the physical properties of rock material can be expected to fluctuate

according to the position of the groundwater table and the moisture content in the rock.
141

The main climatic controls on rock weathering are related to the precipitation,

evaporation and temperature variations within the local environment. The intensity,

frequency and duration of precipitation events, along with seasonal and diurnal

temperature ranges, are important factors in the determination of which physical and/or

chemical weathering processes dominate within a given climatic regime.

Oilier (1984) noted the important influence of topographical attitude of an area on the

weathering processes. Slope angle, whether in a natural slope or engineering structure

such as an embankment or cut slope, can have an important effect because as the angle

increases, weathered products can be more easily removed, thus exposing new materials

to the weathering environment. Relief and slope angle markedly influence the amount

of surface run-off, and thus influence subsurface through-flow of water which affects

the rock materials. The orientation and shape of slopes will help determine the

microclimate, rate of evaporation and soil temperature (Brunsden, 1979).

7.2.2 MINERAL HYDRATION


Mineral hydration causes modifications to the engineering behaviour of soil and rocks

by virtue of changes in volume or density, the interaction between mineral grains, and

the physical properties of the materials involved.

Grim (1962) distinguished two hydration processes in clay soils: namely, intercrystalline

ard intracrystalline swelling. Intercrystalline swelling takes place when the uptake of

moisture is restricted to the external crystal surfaces and void spaces between crystals.

Such swelling may occur in all materials but it is most significant in fine-grained ones,

particularly clays. In relatively dry clays the particles are held together by relict water

under tension from capillary forces. On wetting these forces are relaxed and the

material expands. Such swelling occurs in any type of clay, irrespective of

mineralogical composition, although the amount of swelling depends on a number of

factors including mineral species and the type and concentration of cations present in
142

the porewater. Intracrystalline swelling, on the other hand, is characteristic of the

smectite family of clay minerals, in particular of montmorillonite. Verrniculite and some

varieties of chlorite also display intracrystalline swelling behaviour. In swelling miner

the individual molecular layers of the mineral are weakly bonded so that on wetting

water enters not only between the crystals, but also between the unit layers which

comprise the crystals. Here the magnitude of swelling is a function of clay mineral

type, especially the type of interlayer cations present in the mineral (Taylor and Cripps

1984). For example kaolinite is not expansive whilst montmorillonite is; Na-

montmorillonite being able to expand to many times its original volume. Swelling in

Na-montmorillonite can amount to 800 to 1000 times the original volume (Bell et al,

1986) the clay then having formed a gel of dissociated platelets with dimensions similar

to those of the unit cell (10A). Since swelling is principally due to the ingress of

water, the rock must be porous or fractured. If a rock has an intact unconfined

compressive strength exceeding 40 MPa, it is not subject to swelling (Bell et al, 1986).

Failure of consolidated and poorly cemented rocks occurs during saturation when the

swelling pressure developed by capillary suction exceeds their tensile strength.

7.2.3 MINERAL SOLUTION

Surface and near surface environments can represent conditions in which certain rock

and soil forming minerals are susceptible to chemical change within the context of

weathering. The most important processes, namely solution, oxidation and hydrolysis,

may be controlled by the movement and composition of groundwater since it will act

as the medium of transfer, both into and out of the reaction site, of active components

and products respectively. It is pertinent that the removal of individual grains in

heterogeneous materials will lead to reductions in density and strength, together with

increases in porosity and permeability. In dry air, rocks decay very slowly. The

presence of moisture hastens the rate tremendously, firstly because water is itself an
143

effective agent of weathering and secondly because it holds in solution substances which

react with the component minerals of the rock. The most important of these substances

are free oxygen, carbon dioxide, organic acids and acids of nitrogen.

Weathering of silicate minerals, for example feldspar and mica, is primarily a processes

of hydrolysis. The process whereby feldspars are decomposed to form clay minerals is

affected by the hydrolysing action of weakly carbonated water. Clays are hydrated

aluminium silicates and w h e n they are subject to severe chemical weathering in seasonal

tropical regimes some of them break d o w n to form laterite or bauxite.

7.2.4 PROCESSES AND MECHANISMS OF WEATHERING IN THE STUDY


AREA
It is often difficult to distinguish between physical and chemical processes in rock

weathering because they frequently work together to complement each other.

Nonetheless, for the rocks considered here the physical processes are dominant and

chemical changes are generally only significant in the later stages of weathering, where

the claystone, for example, has been altered to a residual soil. Other than wetting and

drying, physical weathering m a y involve unloading due to erosion (differential release

of confining pressures), which leads to differential stresses and strains in interbedded

sedimentary strata. Considering the effects of weathering on the rock slope stability of

the upper part of the niawarra escarpment, only the earlier stages of alteration are

usually relevant. In this area, the rock is only slightly to moderately weathered, as

failures often occur before the rock has been completely altered to a residual soil.

The fissile claystones are generally m u c h weaker than the more massive claystones.

This is partly due to the ease of moisture movement into the fissile claystone. The

influence of primary sedimentary structures on the breakdown of the sandstones is small.

Clearly the sandstone is m u c h more permeable than the claystone and hence moisture

migration is more important as a weathering factor in the sandstone. The structural


144

changes accompanying an increase in weathering are most important in decreasing the

strength of the rock mass. The effect of wetting (producing swelling) and drying

(desiccation and shrinkage), as shown by the slake durability test, is more pronounced

in some of the claystones, even if the strata have been weathered only slightly.

According to the petrological study (XRD), the Wombarra Shale contains expandable

illite-montmorillonite mixed-layer clay minerals (chapter 4.5.3.4). These minerals expan

by adding interlayer water molecules into their structure and although the process is

reversible, the rocks break down by the generation of unequal local pressures between

grains. Consequently, wet-dry cycling would cause the rocks to disintegrate rapidly as

they do when they are exposed in the vadose zone weathering. Attewell and Farmer

(1976) have discussed the factors influencing the magnitude of the swelling of a clay

in the presence of water. According to their discussion minor sweUing may also

accompany a decrease in confining stress level due to unloading.

Failure of the mineral skeleton along the weakest plane ensues from weathering and an

increased surface area is then exposed to a further sequence of weathering events.

Considering the extremes of rainfall followed by dry periods these cycles are very

common in the Illawarra area. The crack density patterns in claystone illustrate the

increased surface area which this physical breakdown produces. Intra-particle swelling

of the expandable clay minerals is also believed to be very important in the breakdown

of the claystones. As weathering progresses the crack pattern produced by intra-particle

swelling increases the surface area and water intake and drainage is more rapid. Slaking

produced by wetting and drying mechanisms tends to destroy any primary sedimentary

structures, principally bedding. The fissile claystones, even in the fresh state, are mo

open to this type of attack than the massive claystones.

Dearman (1976) has stressed the importance of direct solution in the weathering of

silicate rock material. The solution of siderite and calcite cement in both the sandston

and claystone in the Illawarra area is a contributing factor affecting breakdown of the
145

strata. This solution is usually a slow change, m u c h slower than the associated physical

processes; however, over long periods of time it m a y be more important. Breakdown

of the sandstone and some siltstone is primarily controlled by geological structures

which allow the entry of water. The surface weathering of sandstones can result in

heavy iron staining and case hardening due to the conversion of ferrous oxide to ferric

oxide. The weathering of the sandstone and claystone in a non-oxidising environment,

which is substantially below the watertable, is often quite different. In the latter case,

the claystone and sandstone are permanently saturated and the rate of weathering is

m u c h slower. Solution appears, from thin section studies, to be the major factor

affecting breakdown in this regime but ionic dispersion is also very likely. Chemical

alteration of the component minerals is more evident as the physical processes are

suppressed. The claystone has a more open structure with quite c o m m o n evidence of

solution. The same is seen in the sandstone although secondary deposition of quartz

m a y occur. The siderite and calcite cement are generally reduced in amount and the

shearing strength of the remaining cement and matrix is reduced by the mineralogical

changes resulting from near permanent saturation. However, the rate of strength

reduction is m u c h slower than it is for surface weathering processes, especially wetting

and drying. It is quite possible that salt weathering is also an important breakdown

mechanism for surface rocks in the northern Illawarra area. The salts are provided from

spray and the splashing of waves. Violent storms in the niawarra area are k n o w n to

blow sea spray to great heights and it is quite likely that the entire escarpment would

be covered by sea spray, even the top of the plateau, although the zone just above the

reach of the waves would be most affected. However these effects must be minor in

this area since the vegetation is not specifically salt tolerant. Wellman and Wilson

(1965) listed the conditions necessary for salt weathering: supply of salts; site protected

from wind and rain in which salts can accumulate; and cyclic changes in humidity

and/or temperatures that include the crystallisation point of at least one of the salts
146

present. The salts crystallise in the interstices or pores of the rock and fractures occur

if the stress produced by the growing crystal is greater than the mechanical strength of

the rock. This process is more prone to attacking sandstones than claystones, although

original fissures within claystone could well be opened and expanded by salt weathering.

A slightly weathered sandstone with some outer-most pores exposed would be very

susceptible and may result in honeycomb weathering (e.g. Scarborough Sandstone, south

of Coalcliff along Lawrence Hargrave Drive).

The deposition of salts in the outer pores of rocks could well tend to break the outer

shell of the rock. Wetting and drying, and temperature and humidity fluctuations cause

alternating solution and crystallisation of the salt and create shear stresses in this out

zone which could induce failure. The common observation of so-called "onion skin"

weathering may be caused partially by this mechanism.

Thermal stresses may develop in rocks depend on the rate of heating and cooling.

Although it may be important in some localised situations it is not considered to be an

overall significant physical weathering process in the Narrabeen Group strata, as the

normal range of temperature variation is not very great.

Wind erosion is believed to aid salt weathering and the removal of weathering materials

and so permit deeper weathering; however by itself wind action is not considered to be

a significant cause of breakdown.

The Narrabeen Group claystones were essentially formed in a fresh water environment

and connate water would have had a low salt content. Salts from sea spray find their

way into the groundwater in the Illawarra area. Wallis and Johnson (1969) stated that

Na+ is the principal cationic contaminant in the southern Sydney Basin and CI" is the

main anionic contaminant, thus illustrating the importance of NaCl from sea spray.

These salts, especially Na\ could replace some exchangeable cations in the clays

(principally Ca~ and H+) and there by cause expansion.


147

This is likely to occur more in the massive claystone where randomly orientated clays

were flocculated due to an excess of C a ~ and H + . In contrast, the fissile dispersed

claystones were possibly formed in a more alkaline environment due to excess Na + .

However, this cationic substitution effect would be very small and could almost be

disregarded as a significant breakdown mechanism.

The excess sodium in the groundwater prompted B o w m a n (1972) to believe that sodium

montmoriUonite would tend to form on weathering in preference to calcium

montmoriUonite. It is well k n o w n that the former is much more susceptible to swelling

and breakdown than the latter.

Results of X-ray diffraction analysis indicate that Narrabeen Group rocks and talus

consist of quartz, kaolinite, illite, smectite, mixed-layer clay minerals and iron oxide.

Quartz and kaolinite are abundant in most rocks and talus in the Illawarra area. A n

increase in weathering correlates with an increase in the expandable lattice mixed-layer

clay minerals in the sandstone, shale and claystone. Weathered Stanwell Park Claystone

and W o m b a r r a Shale show a decrease in the kaolinite content and an increase in the

mixed-layer clay minerals.

B o w m a n (1972) recorded analyses of clays resulting from weathering of the Narrabeen

sandstones, with the kaolinite content increasing up the sequence while the content of

mixed-layer clays and illite decrease on ascending stratigraphically. B o w m a n believed

that the illite is degraded into mixed-layer clays with an increase in weathering.

Nonetheless, from aU of the results it is clear that although chemical weathering does

occur, especially in the more highly weathered materials, the main mechanisms of

breakdown of both the claystones and sandstones are fundamentally physical weathering

processes.
148
7.2.5 WEATHERING, STRENGTH AND LANDSLIDES
Weathering of the Narrabeen Group rocks and talus materials will lead to a decrease in

strength which will increase the possibility of landsliding.

(1) The degree of weathering is usually at a m a x i m u m at the ground surface and,

therefore, changes of strength are m a x i m u m near the ground surface. For example,

along the coast line especially between Clifton and Coalcliff.

(2) For sandstone units in the Narrabeen Group, the rate of structural disintegration

and chemical alteration is usually slow so that strength changes of these rocks due

to weathering only need be taken into consideration in the long-term situation.

(3) For claystone and shale in the Narrabeen Group, the influence of structural

disintegration on shear strength is m u c h larger than that which could be normally

caused by chemical alteration of the minerals.

(4) Structural disintegration by weathering occurs more readily and rapidly in claystone

and shale in the Narrabeen Group, which have the capacity for significantly greater

volume change than in the associated sandstone units. T h e latter have little

capacity for volume change.

(5) Structural disintegration occurs most readily in the Stanwell Park Claystone,

W o m b a r r a Shale and the grey claystone interbeds in the Bulgo Sandstone which

fail during minor deformation.

(6) Structural disintegration should also occur very readily in claystone interbedded in

the Coal Cliff Sandstone, which has a great latent capacity for volume change due

to clay minerals (smectite).

(7) Structural disintegration of argillaceous rocks, especially the Stanwell Park

Claystone and claystone in the Bulgo Sandstone in the Illawarra area, can occur

rapidly after either artificial or natural exposure and its effect on rock strength

should be taken into account. O n the other hand, the rate of change of mineralogy
149

is slow, but is important during the long term weathering when considering the

secondary processes of land-sliding.

(8) The effect of chemical alteration (either of mineralogy or matrix material) on the

effective stress and shear strength of talus soils is small in the short-term, to the

point of being inconsequential, when considering the initiation of land-sliding.

(9) The strength of talus soils changes very little due to weathering effects over

human life spans, and landsliding of these materials is primarily initiated by

changes of water pressure, environmental changes, changes of slope geometry,

external loading, excavation, or by such natural occurrences as strong-motion

earthquakes.

7.3 SLAKE DURABILITY TEST


7.3.1 INTRODUCTION
Of particular engineering interest in rocks, especially shales and other mudrocks, is their

durability to weathering upon exposure to surface and underground conditions following

excavation. Weathering effects at the surface are important for assessing the stability

of slopes in road, railway, canal, spillway and other cuts (Dearman, 1976; Rodrigues and

Jeremias, 1990; Hawkins and Pinches, 1992). Durability may vary considerably from

bed to bed in a single outcrop or road cut.

The potential breakdown during slake-durability testing may result from one or more of

the following causes or mechanisms: (a) permeability and porosity which control the

entry and retention of water and its mobility inside the rock; and (b) the action of water

can cause solution of cement, disruption of interparticle bonds, or may set up disruptive

forces due to pore-pressure. Hence a rock that is impermeable usually will be durable.

Clay-bearing rocks, not only mudstone but some sandstone and weathered igneous rocks,

are most susceptible to slake deterioration. The types of clay minerals present are also

important. Sodium clays are easier to disperse than potassium, magnesium and barium
150

clays. In the Illawarra area, for example the dominant clays are illite and

montmorillonite which contain inter-layer cations that favour hydration. In these cases,

swelling of the crystal lattice may well assist in dispersion and disruption processes.

Stress relief is probably also an important mechanism, since consolidated clay-bearing

rocks that have been subjected to burial, tectonic or diagenetic forces are likely to store

elastic strain which will be released if intergranular bonds are weakened by the action

of water.

Various testing procedures are available for the characterisation of argillaceous materials

but their range of applicability, given the need for a useful test result, are significantl

limited. There are many standardised tests for design parameters of soil-like materials,

and conventional rock testing techniques can be applied to heavily compacted or

cemented soils and rocks. There are, however, few test which sensitively quantify the

properties of materials spanning the range between soil and rock.

The major problem in slope stability caused by argillaceous materials is their

susceptibility to degradation upon exposure. Potential results of such behaviour are

rapid slope degradation by loss of strength of the surface material, undercutting of

overlying more competent units, and long-term loss of intact strength affecting the

stability of rock walls.

In the slake durability test (ISRM, 1985) lumps of rock are agitated in a cylindrical

mesh drum while immersed in water. Following this, the material retained in the drum

is oven-dried and weighed. This cycle is repeated; the slake durability index is the

amount of rock remaining after the second cycle, expressed as a percentage of the

original amount. The processes acting in the test are equivalent to those operating

during natural surface exposure. The results may therefore aid in predicting

susceptibility to and rates of surface weathering. It may also be argued that, since the

sample is subjected to stress because of the slaking and abrasion, the slake durability

index may give an indication of rock behaviour under stress conditions. In other words,
151

it m a y be used as a rock index test capable of predicting certain aspects of engineering

performance. However, since the test is rapid, it does not take into account any longer

term dissolution or alteration effects.

7.3.2 SLAKE DURABILITY

The two-cycle slake durability test is an accelerated weathering test. Durability is

defined in this circumstance as the resistance of a rock to weakening processes. It is

really the inverse of the term weatherability. The durability index can vary from 0 %

where a material completely disintegrates, to 1 0 0 % durability w h e n no disintegration

takes place.

A brief review of the likely processes of degradation operating in the slake durability

test is necessary to establish the applicability of the test results. The test is a

combination of slaking (i.e. breakdown upon exposure to moisture) and abrasion. V a n

Eeckhout (1976) concluded that shear strength was reduced by expansion of fractures

due to capillary tension changes, pore pressure increase, friction reduction and chemical

deterioration. If these mechanisms of strength reduction (rather than active breakdown)

are relevant in the slake durability test, then the abrasive action of the test would be

significant. However, Hudec (1976) found that losses in the slake durability test were

no more than those in an alternating wet-dry test. Thus it seems that the slaking effect

is the most important process; the agitation merely enables all the small fragments

generated to pass through the mesh.

The mechanisms causing slaking breakdown are, however, far from completely

understood. Various authors have used liquids of different surface tensions to elucidate

the slaking process.

Colback and Wiid (1965), Nakano (1967), Taylor and Spears (1970) all recognised a

decrease in durability or strength when rocks were immersed in liquids of increasing

surface tension. Oliver (1980) summarised work based on strain measurement of shale
152

during changes in humidity and moisture content. The degrading processes were related

to partially irreversible anisotropic expansion and shrinkage of the rock w h e n subjected

respectively to capillary action and drying. Also noted was the relative importance of

microfracturing in controlling breakdown compared with the influence of mineralogy.

These microdiscontinuities provide the conduits for the moisture redistribution that

resulted in slaking.

7.3.3 AIM OF STUDY

The primary objectives of this study were to: (a) identify the durability index and

delineate the engineering properties of strata in the lower Narrabeen Group. Such index

properties would be applied in a classification scheme for engineering purposes and

possibly to aid in geological classification; (b) derive a relationship between the slake

durability result and the mineralogy and fabric of the rocks; (c) account for differing

behaviours of shale units of similar age; and (d) interpret the processes occurring in the

test.

7.3.4 METHOD OF STUDY

The method of study can be divided into subsections as follows.

7.3.4.1 SAMPLE COLLECTION

Samples of weathered rocks were collected from the northern niawarra district between

Clifton and Coalcliff, and between Scarborough Station and Stanwell Park Station. Coal

Cliff Sandstone samples came from the base of the landslide at Coalcliff Harbour.

Wombarra Shale, Otford Sandstone M e m b e r and Scarborough Sandstone samples came

from cliffs beside the road near the Jetty Fault and the railway cutting at Scarborough

Station. Bulgo Sandstone samples came from the railway cutting at Stanwell Park

Station,
153

Core samples were collected from boreholes IL64 and IL55, drilled in the North Cliff

area (Fig. 3.11) approximately 10 km west of the coastal exposure.

7.3.4.2 SAMPLE PREPARATION

Samples from each site consisted of ten representative, intact, roughly equidimensional

rock fragments weighing 40 g to 60 g each (ASTM, 1991). These fragments were

produced by breaking the rock with a hammer. Fragments were obtained from rock

cores and from weathered rock outcrops. Sharp corners were broken off and dust was

removed by brushing the sample just prior to weighing. The total sample weight from

each site was 450 g to 550 g.

7.3.4.3 PROCEDURE

(1) For each test the rock fragments were placed in a mesh drum. They were

weighed, and dried in an oven (110C) for 16 h. The rock and drum were

allowed to cool to room temperature for 20 minutes and weighed again. The

natural water content was calculated as follows:

W = [(A - B)/(B - Q] x 100

Where

W = percentage water content

A = mass of drum plus sample at natural moisture content

B = mass of drum plus oven-dried sample before the first cycle

C = mass of drum

(2) The drum was mounted in the trough and coupled to the motor. The trough was

filled with tap water, at room temperature, to 20 mm below the mesh drum axis.

The mesh drum is rotated at 20 rpm for a period of 10 minutes.


154

(3) The mesh drum was removed from the trough immediately after the rotation period

was completed and the mesh drum and the retained sample were dried in the oven

for a further 16 h.

The mesh drum and sample were weighed to obtain the oven-dried mass for the

second cycle. Steps 2 and 3 were repeated. Again the drum and sample were

weighed to obtain a final mass.

7.3.4.4 CALCULATIONS

The slake durability index is calculated, as follows :

Id2 = t(Wf - C)/(B - C)] x 100

Where

Id2 = slake durability index (second cycle)

B = mass of mesh drum plus oven-dried sample before the first cycle

W f = mass of mesh drum plus oven-dried sample retained after the second cycle

C = mass of mesh drum.

7.3.5 RESULTS

Thirty six samples were subjected to slake durability testing. The samples ranged from

fresh to highly weathered claystone, shale, coal, sandstone and interbedded claystone.

A third and fourth cycle was performed in an effort to make the test more realistic for

long term weathering (Figs 7.1 to 7.8). The results are given in Tables 7.1 and 7.3.

The slaking fluid was tap water at 21C. The slake durability index is defined after the

second cycle and it is used for slake durability classification. According to Hopkins and

Deen (1984) the in-situ (natural) moisture content of shales provides a strong indication

of their slake-durability properties. They considered shales with a natural water content

below approximately 3.5% to have a high slake-durability index while those with natural

water contents of between 3.5% and 7.5% appear to have an intermediate slake-
155

durability rating. In general, there is a decrease in durability from the fresh to

weathered specimens, especially for the Stanwell Park Claystone (Figs 7.4b, 7.5a). The

durability of this claystone is highly dependent on its water content (Fig 7.8a), the

durability decreases rapidly as water content increases. Grey claystone interbedded in

the Bulgo Sandstone has a very low durability (Figs 7.6 and 3.9a) and appears to be

akin to the underlying Stanwell Park Claystone. Weathered Wombarra Shale has a

natural water content between 1.43% and 7.25% (Fig 7.8b), and the durability decreases

as water content increases (samples were collected one week after heavy rainfall).

Sandstone samples 1, 2, 3, 12, 13, 14, 15 (Fig 7.9), 16, 17, 18, 19, 25, 26 and 27, and

shale samples 17, 18, 19, 22 and 23 showed only relatively a small breakdown, whereas

claystone samples 4 and 5 showed a moderate durability. Claystone samples 6, 7, 8,

9 and 11 showed lower slake durability index values for weathered rocks compared to

fresh claystone rocks, especially samples 4 and 5 (Fig 7.10) which belong to the Stawell

Park Claystone.

Field observations and slake durability tests indicate a distinct difference between the

rates of weathering of the claystone interbedded in the Narrabeen Group sandstone. In

(Table 7.2) grey claystone interbedded in the Bulgo Sandstone (Fig. 3.9a; samples 1 and

2) showed lower slake durability, whereas claystone interbedded in the Scarborough

Sandstone (Fig. 3.6; samples 3 and 4) and claystone interbedded in the Coal Cliff

Sandstone (Fig. 3.22; samples 5 and 6) showed medium slake durability and only a

small breakdown respectively. It is dependent on their mineralogy, and especially on

the amount of carbonate cement. Distinct differences in durability between the coal

(sample 3 in Table 7.3) and highly weathered sandstone (samples 1, 2 in Table 7.3) in

the niawarra Coal Measures may have important consequences for slope stability in this

formation in the Clifton area.


156

7.3.6 SLAKE DURABILITY CLASSIFICATION

Gamble (1971) proposed using the results from the second 10 minute cycle after drying

as a basis for slake durability classification. Values of the slake durability index for

representative shales and claystones tested by Gamble varied over the whole range from

0% to 100%. Based upon his results, Gamble proposed a classification of slake

durability (Table 7.4). Franklin and Chandra (1972) proposed a different subdivision

for slake durability classification (Table 7.5). Slake durability classification for

Narrabeen Group strata according to Franklin and Chandra (1972) is shown in Table 7.6.

7.3.7 STATIC (LONG-TERM) DURABILITY TESTING

Static durability tests were carried out on selected samples from the Narrabeen Group

which included both fresh core and weathered outcrop materials. Each sample was

subdivided into two portions, which were weighed and then tested with distilled water

or tap water respectively. Each sample was placed in a beaker, that was filled with

water to provide about 20 mm sample cover. Visual assessments of the specimen

condition were made at elapsed times of 1 minute, 10 minutes, 1 hour, 5 hours,

24 hours (Fig. 7.11) and two days. The specimen condition was reported as a

letter/number code as defined in Table 7.7.

In an effort to make the tests more realistic for long term weathering, the samples were

subjected to repeated wetting and drying events to simulate wet and dry conditions on

the outcrop. The samples were, therefore, drained and dried in an oven (65C) for 24

hours. The rock and beaker were allowed to cool to room temperature for 10 minutes

and weighed again. This process was then repeated after one week, two weeks

(Fig. 7.12), four weeks and then at monthly interval to six months or until the sample

totally disintegrated. The results are shown in (Tables 7.8 to 7.13) for fresh and

weathered samples respectively. Solubility and disaggregation of the Coal Cliff

Sandstone, Otford Sandstone Member, Scarborough Sandstone and Bulgo Sandstone in


157

water are shown in Figures 7.13 to 7.15 and Table 7.14. During the 6 months of

testing the percentage of weight loss per cycle for the Coal Cliff Sandstone,

Scarborough Sandstone and Bulgo Sandstone were 0.3%, 0.17%, and 1.11% respectively.

During the first two months the percentage of weight loss per cycle for the Otford

Sandstone Member was 1.22% but at the end of two months the sample became

completely disaggregated.

7.3.8 CONCLUSIONS

Slake durability tests approximately reproduce, on an experimental scale, the conditions

in the cliffs that cause differential erosion. They illustrate an accelerated reduction in

durability from repeated wetting and drying, which is one of the main factors affecting

the surface layers of the cliffs (Norris, 1990). Also it is useful for comparing the

weathering characteristics and associated lithological changes between the Wombarra

Shale and Stanwell Park Claystone and claystone interbedded in the Narrabeen Group

sandstone. Moderately and highly weathered Stanwell Park Claystone samples have very

low durability (Fig. 3.8); it is dependent on their mineralogy, and especially on the type

and quantity of clay minerals present. Different clay minerals have different influences

on the mechanical behaviour of rock. Cripps and Taylor (1981) demonstrated the

importance of expansive clay minerals within argillaceous rocks in controlling moisture

sensitivity. The predominant lithological characteristic of claystone is its high proportion

of detrital and authigenic clay (Dick and Shakoor, 1992). Consequently, the clay

minerals have the most pronounced influence on the durability behaviour of claystone.

Kaolinite, illite, mixed-layer (illite-smectite) and montmorillonite are common in the

Stanwell Park Claystone and grey claystone interbedded in Bulgo Sandstone. The

expandable mixed-layer illite-smectite and montmorillonite varieties are particularly

important because their presence makes the Stanwell Park Claystone highly susceptible

to slaking when exposed to water. The relatively high durability of the Wombarra Shale
158

is attributed to its higher degree of consolidation and cementation compared with the

Stanwell Park Claystone.

It is the structural laminations in shale that distinguishes it from other mudrocks. The

laminations are, in part, a manifestation of the degree of uniform orientation of platy

minerals in the rock (Potter et al, 1979). A s the degree of orientation increases, so

does the degree of consolidation.

Field observations show that the Wombarra Shale is well laminated in parts of formation

indicating a relatively high degree of consolidation.

Presumably the main problem with the slake durability test is that it is not totally

analogous to the natural processes which are acting on the cliffs. It gives quantitative

data concerning the rocks but, unless the complex processes of nature can be reproduced

in the laboratory, the results can not be expected to necessarily give values which are

directly useable. However, they should give the relative proportion of disintegration that

each lithology is likely to undergo.

The third and fourth cycles were performed in an effort to make the slake durability test

more realistic. However, the results are not significantly different from the previous two

cycle losses. The extended testing was useful for the Stanwell Park Claystone and

weathered W o m b a r r a Shale samples because they showed m u c h lower index values.

Although the tests show little difference in durability or strength, except for the Stanwell

Park Claystone ( M W and H W samples), the lithologies are in fact different for each

formation, and the differential erosion in the cliff sections is probably controlled by

these differences.

The Coal Cliff Sandstone has a clay matrix content of around 3 0 % of the total rock.

It consists of cross-bedded sandstone with interbeds of shale and sandy shale. Marine

abrasion attacks these sandstones just above mean sea level at the base of the cliff and

tends to wash the clay from between the quartz grains in the sandstone beds. This

would cause a large increase in the porosity and also a significant decrease in the
159

strength of the surface layers. Small rock fragments or individual crystals can then fall

off, or be washed off by waves. The increased porosity of the surface would also

facilitate the deposition of salt from evaporating sea water. Such salt deposits expand

during crystallisation and can force off grains or thin layers of the surface material by

repeated solution and precipitation mainly just above high tide level.

In the Scarborough Sandstone and Bulgo Sandstone, the quartz content is higher, and

the rocks appear to be strongly cemented with calcite and kaolinite. If the clay were

washed out, the porosity increase would be less than the increase suffered by the Coal

Cliff Sandstone, so that the strength decrease of the surface layers would be marginal.

It has been well established that the moisture content in argillaceous rock masses such

as claystone and shale can have a significant effect on their properties. The

compressive strength of quartzitic shale under saturated conditions is in the order of

5 0 % of that under dry conditions (Colback and Wild, 1965). The weathered Stanwell

Park Claystone has a high water content which m a y be a useful index for determining

other engineering properties, especially the strength of the intact rock.

The petrological study (see chapter 4) also provided a useful indication of the character

of the rock, and gave a guide to the likelihood of differential erosion. The latter is

dependent on the clay content of the sandstone, and this feature, combined with the

secondary cement, is the most important factor controlling the erosion of these coastal

cliffs.

Sandstone units in the Narrabeen Group, which contain abundant clay minerals and

volcanic rock fragments, show significant strength loss on wetting. In the clay-rich

varieties the change in strength is likely to be related to the softening and possible

expansion of the clay minerals. The petrological study in this research showed that the

Narrabeen Group sandstone contains swelling clays. Therefore, expansive forces in these

rocks are a mechanism contributing to strength loss. High proportions of expansive clay

minerals were detected in volcanic rock fragments (cherts) which suggest that clay
160

softening in the presence of water is important in controlling moisture related reductio

in strength of sandstone in the Illawarra area. The nature of the cement also influences

slake durability. Sandstone units in the Narrabeen Group containing kaolinite and calcit

cements are generally less susceptible to moisture effects. Therefore, slake durability

of intact samples from the Narrabeen Group rocks is controlled primarily by the

mineralogy of rocks and to a lesser extent by the rock microfractures.

The static durability test was performed in an effort to make the durability more

realistic. The results are useful for a better understanding of weathering and slope

stability. In the long term, solubility and disaggregation in the Otford Sandstone

Member would cause an increase in the secondary porosity of this sandstone. The

percentage of weight loss per cycle for the Otford Sandstone Member was 1.22% and

at the end of two month the sample become completely disaggregated. Small rock

fragments or individual crystals fell off and in the outcrop situation could have been

washed off by surface or groundwater movement (Fig. 3.5). The percentage of weight

loss per cycle for the Coal Cliff Sandstone, Scarborough Sandstone and Bulgo Sandstone

are 0.3%, 0.17%, and 1.11% respectively. The results of solubility and disaggregation

provide answers as to why the cliffs in Narrabeen Group strata have high fracture

porosity. It would appear that zones with soluble cement are eroded more quickly than

adjacent areas with less soluble material.

7.4 POINT LOAD STRENGTH TEST


7.4.1 INTRODUCTION

Strength determination of a rock usually requires, careful test set up and specimen

preparation, and the results are highly sensitive to the method and style of loading.

Many methods are available for the strength determination of rocks. An easy and

inexpensive field technique for measurement of rock strength is the point load test

described by Broch and Franklin (1972). This test provides an index for the strength
161

classification of rock materials. It may also be used to predict other strength parameters

with which it is correlated, for example uniaxial tensile and compressive strengths

(ISRM, 1985).

7.4.2 AIM OF STUDY


Point load strength tests were carried out so that: (a) a greater number of weathered

specimens could be tested for comparison with test results from fresh specimens; (b)

tests could be made both perpendicular and parallel to bedding and hence a

consideration of effect of anisotropy could be made; and (c) the strength classification

of the various rock units could be determined.

7.4.3 METHOD OF STUDY


The method of study can be divided into seven sub-sections as follows.

7.4.3.1 SAMPLE COLLECTION AND PREPARATION


Approximately 240 samples were gathered for point load testing.

Samples of weathered rocks (irregular lumps) were collected from the northern Illawarra

district between Clifton and Coalcliff, and between Scarborough Station and Stanwell

Park Station. Coal Cliff Sandstone samples were obtained from the base of the

landslide at Coalcliff Harbour. Wombarra Shale, Otford Sandstone Member and

Scarborough Sandstone samples came from cliffs beside the road near the Jetty Fault

and from the railway cutting at Scarborough Station. Bulgo Sandstone samples came

from the railway cutting at Stanwell Park Station.

Core samples were collected from Kembla Coal & Coke (KCC) boreholes IL64 and

IL55, drilled in the North Cliff area (Fig. 3.11). Diameters of core samples were 60

mm and diameters of irregular lumps were generally greater than 40 mm but this was

not always possible. The International Society for Rock Mechanics (1972) has
162

suggested a minimum of twenty lumps for the irregular lump test; this was possib

to the large number of suitably weathered outcrop samples.

7.4.5 DIAMETRAL TESTS

Core samples with a length/diameter ratio greater than 1.0 were used for diametr

testing. At least 10 tests were carried out per sample. Each sample was inserted

test machine and the platens closed to make contact along a core diameter. The

distance L between the contact points and the nearest free end was at least 0.5

the core diameter (Fig 7.16a). The load was steadily increased so that failure oc

within 1-2 minutes, and the failure load P was recorded. If the fracture surface

through only one loading point (Fig 7.17c) the test was carried out again.

7.4.6 AXIAL TESTS

Core samples with a length/diameter ratio of 0.3-10 were used for axial testing

7.16b). At least 10 tests were carried out per sample. The sample was inserted in

test machine and platens closed to make contact along a line perpendicular to th

end faces. The load was steadily increased so that failure occurred within 1-2 m

and the failure load P was recorded. If the fracture surface passed through only

loading point (Fig. 7.17d), the test was carried out again.

7.4.7 IRREGULAR LUMP TESTS

Irregular lumps with a ratio D/W between 0.3 and 1.0 were used for these tests (

7.16c). At least 10 tests were carried out per sample. The specimen was inserted

the testing machine and the planes closed to make contact with smallest dimensio

the lump.
163

The load was steadily increased so that failure occurred within 1-2 minutes, and the

failure load P was recorded. If the fracture surface passed through only one loading

point, the test was carried out again.

7.4.8 CALCULATIONS

The point load index (L) indicates the rock strength at the time of failure;

Is = P/D2

The uncorrected Point Load Strength L. was calculated as P/De2 where De, the "equivalent

core diameter", was given by:

De2 = D2 for diametral tests

De2 = 4A/ for axial block and lump tests

A = WxD = minimum cross-section area of a plane through the platen contact points

(Fig. 7.16c).

Since Is varies as a function of D in the diametral test, and as a function of De in axial,

block and irregular lump tests, a size correction must be applied to obtain a unique

Point Load Strength value for the rock sample. The latter can be used for the purpose

of rock strength classification.

The size-corrected Point Load Strength Index L(50) of a rock sample is defined as the

value of Is that would have been measured by a diametral test with D = 50 mm.

The "Size Correction Factor F" can be obtain from the expression:

F = (D/50)045 (after Greminger, 1982)

For general purposes and for tests near the standard 50 mm size, very little error was

introduced by using the approximate expression:

F = 0V50) (after ISRM, 1985)

Therefore the necessary formulae for the strength calculation could be written as :

L = F.P/De2 (after ISRM, 1985)


164

M e a n values of Is(50) can be used when classifying samples with regard to their Point

Load Strength anisotropy indices. The mean value was calculated by deleting the

highest and lowest values from the conducted valid tests and averaging the remaining

values.

The strength anisotropy index 1.(50) was defined as the ratio of the mean 1,(50) values

measured perpendicular and parallel to planes of weakness. This ratio would be near

1.0 for quasi-isotropic rocks and higher values would be obtained for anisotropic rocks.

7.4.9 RESULTS

The Point Load Strength classification for Narrabeen Group samples is shown in Table

7.15 for fresh core samples, and Table 7.16 for weathered samples. Summarised point

load strength results for the Coal Cliff Sandstone, Wombarra Shale, Otford Sandstone

Member, Scarborough Sandstone, Stanwell Park Claystone and Bulgo Sandstone are

shown in Tables 7.17 to 7.22. In general, there is a decrease in strength from the

fresh to weathered specimens and the strengths parallel to the bedding are generally

much lower than those normal to bedding (Fig. 7.18). The fresh core samples are very

strong (Table 7.15); in contrast, some of the weathered specimens sampled at the surface

show low strength as noted for moderately weathered Stanwell Park Claystone (Table

7.16 and Fig. 7.19). The Ia(50) results indicate the effect of orientation on the strength

(Fig. 7.20). The sandstones are generally only slightly anisotropic, whereas fresh shale

and claystone are strongly anisotropic.

M a n y surface samples show medium to high strength, for example highly weathered

Bulgo Sandstone, highly weathered Scarborough Sandstone and moderately weathered

Wombarra Shale (Table 7.16). S o m e of the weathered samples are still strong to very

strong, for example moderately weathered Bulgo Sandstone, Scarborough Sandstone and

Coal Cliff Sandstone (Table 7.16). In these cases, it seems likely that the weathering

of the samples has caused alteration of the siderite cement to hematite and other iron
165
oxides cements, thus producing secondary hardening. A m o n g the weathered specimens,

Otford conglomerate is interesting because it is very strong (probably because it is well

cemented).

7.4.10 RELATIONSHIP BETWEEN POINT LOAD STRENGTH INDEX AND


UNIAXIAL COMPRESSTVE STRENGTH
Various correlations between the Point Load Strength Index (PLSI) and the Uniaxial

Compressive Strength (UCS) have been reported in the literature (e.g. Bieniawski, 1975;

Ghosh and Srivastava, 1991; Tsidzi, 1991). Bieniawski (1975) reported that for a

variety of rock types the global average (UCS) to Is ratio was approximately 23.50.

Ghosh and Srivastava (1991) recently reported that a ratio of 16 was more appropriate

for rocks from India.

For this study, the average Is w a s correlated with the (UCS) of the rock by the

following equation (after Bieniawski, 1975):

U C S = (14 + 0.175 D J L

Based on tests concluded in the laboratory on core and lump samples the relationship

between PLSI and U C S for various members of the Narrabeen Group are presented in

Tables 7.17 to 7.22.

In general there is a decrease in strength from the fresh to weathered specimens (Figs

7.21 to 7.23). The fresh samples are strong to very strong; in contrast, some of the

weathered specimens sampled at the surface are moderately strong to strong (Table

7.23). A m o n g the weathered specimens, moderately weathered Stanwell Park Claystone

is interesting because it is moderately weak (probably because of mineral composition).

The strengths parallel to the bedding are generally much lower than the strength

perpendicular to bedding (Fig 7.24). This reflects the presence of oriented elongate and

platy grains or clearly defined bedding planes.


166

7.4.11 CONCLUSIONS

The majority of samples subjected to the point load testing were sandstone because

much of the Narrabeen Group consists of sandstone and the shale units are commonly

unsuitable for testing because of the presence of closely spaced joints. The weathered

Stanwell Park Claystone and weathered Wombarra Shale usually broke up prior to

testing, or at the start of testing the specimens broke immediately upon the application

of the load before obtaining any record on the pressure gauge. It was not possible to

test highly weathered claystone at all.

Anisotropy seriously affected the test values. Failures very often occurred along the

bedding planes rather than through the rock substance, particularly when the rock was

weathered.

These results provide answers as to why the cliffs in Narrabeen Group strata are

undergoing differential erosion. Differential erosion is the result of weathering of rocks

which are not uniform in character, but are softer or more readily altered in some places

than in others. The result is usually an uneven surface, with the softer rocks being

removed more quickly than the harder strata. As expected, the data presented here

show that the fresh rocks are stronger than weathered rocks. The Scarborough

Sandstone and Bulgo Sandstone are marginally stronger than the Coal Cliff Sandstone.

The Wombarra Shale is stronger than the Stanwell Park Claystone. Based on the

Uniaxial Compressive Strength fresh Scarborough Sandstone and Bulgo Sandstone can

be classified as being very strong whereas fresh Coal Cliff Sandstone is strong to very

strong. Fresh Wombarra Shale can be classified as moderately to very strong whereas

fresh Stanwell Park Claystone is moderately strong to strong. Weathered sandstone in

the Narrabeen Group is moderately strong to strong whereas measured values from

weathered shale and claystone range from moderately weak to moderately strong. The

highly weathered shale and claystone were too weak to test.


167

Strength is related to porosity, amount and type of cement and/or matrix, the

composition of the individual grains, and the amount and type of weathering. The

Narrabeen Group sandstones all contain variable proportion of quartz, clay and carbonate

cements, with the content of the latter generally being lower in weathered samples than

in fresh core material.

The amount of cementing material is more important than the type of cement, although

if two sandstones are equally well cemented, one having a siliceous, the other a

calcareous cement, then former is the stronger (Bell, 1983). Thus the cemented shales

are invariably stronger and more durable than poorly cemented varieties whereas

carbonaceous shale containing a significant proportion of organic matter, is still less

durable. Moderate weathering increases the fissility of shale by partially removing the

cementing agents along the laminations or by expansion due to the hydration of clay

particles. This has certainly occurred in the Wombarra Shale and accounts for the

relatively low strength of moderately weathered Wombarra Shale (Table 7.23).

Petrological studies (chapter 4) provided a useful indication of the character of the

sandstone, and give a reasonable guide to the likelihood of differential strength. The

clay content of the rocks, combined with the secondary cement, is probably the most

important control on the differential strength of these sandstone units.

7.4.12 RELATIONSHIP BETWEEN UNIAXIAL COMPRESSIVE STRENGTH


(UCS) AND SLAKE DURABILITY INDEX (SDI)
The relationship between uniaxial compressive strength (UCS) and slake durability index

(SDI) is shown in Fig. 7.25. The relationship between uniaxial strength (UCS) and

slake durability index (SDI) for weathered Wombarra Shale and weathered Stanwell Park

Claystone are shown in Figs 7.26 and 7.27. In general, there is a decrease in strength

and durability from fresh to weathered specimens. The fresh samples which are strong

to very strong show extremely high durability; in contrast, weathered samples which are
168

moderately strong to strong show high to very high durability. Among the weathered

specimens, moderately and highly weathered Stanwell Park Claystone are moderately

weak to moderately strong and show very low durability.

7.5 ROCK COMPOSITION IN RELATION TO MECHANICAL PROPERTIES


The petrographic description of rocks for engineering purposes includes the

determination of all parameters which have a bearing on the mechanical behaviour of

the rock or rock mass but cannot be obtained from a macroscopic examination of a rock

sample, for example, parameters such as mineral content, grain size and texture

(Hallbauer et al, 1978). Microscopic characteristics, in particular in the areas of grain

contact and cementation, explain the strength and behaviour of sandstones (Dyke and

Dobereiner, 1991).

The ISRM recommends that the report of a petrographic examination should be

confined to a short statement on the origin, classification and details relevant to the

mechanical properties of the rock concerned.

7.5.1 ROCK COMPOSITION AND STRENGTH


The compressive strength of a sandstone is influenced by its porosity, amount and type

of cement and/or matrix material as well as the composition of the individual grains

(Hawkins and McConnell, 1991, 1992; Haney and Shakoor, 1994). Price (1963) showed

that the strength of sandstones with a low porosity (less than 3.5%) was controlled by

their quartz content and degree of compaction. In those sandstones with a porosity in

excess of 6% he found that there was a reasonably linear relationship between dry

compressive strength and porosity, for every 1% increase in porosity the strength

decreased by approximately 4%. If cement binds the grains together then a stronger

rock is produced than one in which a similar amount of detrital matrix performs the

same function. However, as noted previously the amount of cementing material is more
169

important than the type of cement (Bell, 1983). For example, ancient quartzarenites, in

which the voids are almost completely occupied with siliceous material are extremely

strong with crushing strengths exceeding 240 MPa. By contrast poorly cemented

sandstones may possess crushing strengths of less than 3.5 MPa. In the Narrabeen

Group sandstones, porosity is generally less than 3.5 % (see chapter 4) and their

strength was controlled by their quartz content (Tables 7.24 to 7.26). Figures 7.28 to

7.31 show a reasonably linear relationship between unconfined compressive strength and

percentage of quartz for fresh and slightly weathered Coal Cliff Sandstone samples, fresh

Scarborough Sandstone samples, fresh Bulgo Sandstone samples and the total fresh,

slightly and moderately weathered Narrabeen Group sandstones samples respectively.

Also there is a linear relationship between point load strength index and percentage of

quartz for fresh, slightly and moderately weathered Narrabeen Group sandstone samples

(Figs 7.32 and 7.33).

7.5.2 ROCK COMPOSITION AND SLAKE DURABILITY


Mudrocks are sedimentary rocks in which more than 50% of the grains have a diameter

of less than 0.062 mm. Mudrocks can be distinguished on the basis of percent clay and

presence of structural laminations (Potter et al, 1979). Within the mudrock group, five

different classes of rocks are recognised: shale, mudstone, claystone, siltstone and

argillite. It has been estimated that mudrocks account for as much as 70% of all

sedimentary rocks (Picard, 1971). Consequently, mudrocks are frequently encountered

in all types of geotechnical engineering projects. Where mudrocks are excavated and

left exposed to weather, they rapidly slake to produce a soil-like material having

significantly inferior engineering properties to those of the original rocks. In these

situations, the durability of mudrocks becomes their most important engineering property.

The prediction of mudrock durability is complex because there are multiple lithological

factors controlling durability. This complexity is aggravated by the very fine-grained


170

nature of mudrocks which makes the lithology difficult to study except by XRD (Table

7.27). The durability of shales is related to the fabric as expressed by void ratio and

absorption. Mudstones are distinguished on the basis of an absence of well developed

structural laminations and the presence of micro-fractures. The durability of mudstones

is related to the frequency of the micro-fractures. The durability of argillite is not a

problem. The mineral grains in argillite have been recrystallised resulting in a very

durable rock (Russell, 1981). Obviously, the presence of swelling clays in any of these

rocks will have an effect on durability.

Testing of samples from the Wombarra Shale has shown that the slake durability test

is capable of making a sensitive differentiation between shale of varying composition.

It is reasonable to expect that properties such as strength will vary with these

compositional changes. The form and amount of calcite is the principal control over

large scale variations of the slake durability index in the Wombarra Shale for fresh

samples (1 and 2 in Table 7.27) and weathered samples (3 to 8 in Table 7.27). The

presence of calcite in shale will usually cause high durability. The Wombarra Shale is

cemented in the Clifton area beside the Jetty Fault but north of Wombarra Station it

is not. Another difference in the behaviour of samples is that the slake durability of

Wombarra Shale from south of Wombarra Station (samples 7 and 8 in Table 7.27) is

sensitive to the abundance of clay minerals in the clastic fraction whereas the shale in

the Clifton area is not (samples 3 and 4 in Table 7.27). The shale south of Wombarra

Station is less durable than that at Clifton perhaps because of its fabric. Reflecting th

arrangement of individual clay particles, the microcracks, which allow access for water

and along which degradation is concentrated, have greater curvature in the fissile shale

south of Wombarra Station. The cracks will, therefore, tend to meet more frequently

during slaking of this shale, generating smaller and more equant particles which will fal

through the test mesh. There is, therefore, a complex interplay of geological factors tha

control the durability and presumably all the other properties of the Wombarra Shale
171

from south of Wombarra Station and from Clifton. Similar variability is expected in

other units of similar lithology.

A higher proportion of clay minerals in any rock may be expected to cause a greater

tendency to slake. For example pure claystone durability is controlled by clay content.

The complex interaction of sedimentation conditions, diagenetic processes and stress

history that produce claystone apparently causes them to have a generally low durability.

Testing of samples in the Stawell Park Claystone show that this claystone to the north

of Coalcliff Station has a lower durability than that located beside the Harbour Fault

(Table 7.28). Variation in durability within this claystone is again controlled by

mineralogy. The distinct difference between the durability of fresh samples 1 and 2

(Table 7.28) and weathered samples 3 and 4 (Table 7.28) is controlled by the quantity

and condition of the clay minerals. The diffraction intensity of clay minerals in samples

3 and 4 is more than that in samples 1 and 2 (Table 7.28).

Testing of samples of interbedded claystone in the lower Narrabeen Group sandstone

units shows different durability in each unit (Table 7.29).

Claystone interbedded in the Bulgo Sandstone (samples 3 and 4, Table 7.29) shows very

low durability. Claystone interbedded in the Coal Cliff Sandstone (samples 1 and 2,

Table 7.29) have a high durability whereas claystone interbedded in the Scarborough

Sandstone (samples 5 and 6, Table 7.29) shows medium durability. The differences in

the behaviour of the samples is that slake durability is sensitive to the abundance of

clay minerals in these samples. Differences between the durability claystone interbedded

in the Coal Cliff Sandstone (samples 1 and 2, Table 7.29) is related to the intensity of

carbonate cementation. The claystone interbedded in the Bulgo Sandstone and highly

weathered Stanwell Park Claystone have very low durability (samples 3 and 4 in Table

7.29 and sample 11 in Table 7.1). This has a significant effect on slope stability in

the Bulgo Sandstone and in the Stanwell Park Claystone while acts as the bedrock for

much of the talus mantle between Clifton and Stanwell Park.


172

7.5.3 RELATIONSHIPS B E T W E E N SLAKE DURABILITY, R O C K S T R E N G T H

AND WEATHERING IN RELATION TO ROCK SLOPE AND TALUS


FAILURE ALONG THE NORTHERN ILLAWARRA COASTLINE
In general, weathering has a major effect on the strength and slake durability of rocks

in the northern Illawarra region since they consist of unstable lithic sandstone largely

derived from a volcanic source. As the degree of weathering increases, the rock masses

become more porous and weaker with clay minerals concentrated along bedding planes

and joint surfaces. Therefore, a consideration of the weathering state of the rock, and

its resultant strength and durability, is essential for assessing the relative stability of

rock.

The effects of weathering and alteration processes on the weak sedimentary rock

sequences in the niawarra region has been studied here. Weak sedimentary rocks

(shale, siltstone and mudstone) are commonly characterised by having a high sensitivity

to variations in moisture content that cause alternating swelling and shrinking, especially

in smectite-rich volcanic derived sequences. Volume changes within the rock mass

result in significant strength reduction and considerable deformation occurs as a

consequence. These volume changes are very difficult to control without an extensive

drainage system being emplaced both on the surface slope and within the adjacent rock

mass. Most cases of rock-slope instability within the northern Illawarra can generally

be directly attributed to the alteration and weatherability of the weak rock units.

In the argillaceous rocks in the study area, although the differences in lithology and rock

moisture content are small, weathering produce considerable changes in the mechanical

properties of adjacent layers. Weathering is limited in depth to the zone above the

permanent water table position. Above the capillary zone, because of the great variation

in moisture content, weathering processes increase, causing intense fracturing and

strength reduction in the surficial portion of the rock mass.


173

In the low permeability rocks in the northern Illawarra region, a 0.5 - 1 m thick layer

of very weathered rock is formed at the surface by the fluctuation of moisture due to

cycle of wet and dry periods. Apparently this layer acts as a natural protection against

the further development of weathering process, but if removed by slippage or slumping,

the process is reactivated, initiating the development of a n e w layer. Because sandstone

layers are interbedded with the argillaceous rocks, and these low permeability coarser

rocks have a higher strength and resistance to weathering, they form a series of small

benches which are generally fractured along joints, creating unstable blocks.

The rocks in the northern Illawarra area have been subjected to unloading stresses due

to weathering and scarp formation (differential release of confining pressures), which

has led to differential stresses and strains being generated in interbedded sandy and

argillaceous strata. T h e fissile claystone units are generally m u c h weaker than the

associated sandstone strata. This is partly due to the ease of moisture m o v e m e n t into

the fissile claystone resulting from expansion and contraction of clay minerals. Hence

an increase in weathering of the claystone beds is the most important factor controlling

the decrease in strength of these rocks and thus increasing slope instability.

In the Narrabeen Group rocks slake durability is low in claystone laminae interbedded

in the Coal Cliff Sandstone, and its effect on rock strength should be taken into account

even if the adjacent strata have been weathered only slightly. The rate of change in the

physical characteristics of the rock mass due to slaking is slow, but it is important

during longer term weathering, especially when considering the secondary processes of

land-sliding. Slaking produced by natural wetting and drying mechanisms tends to

destroy any primary sedimentary structures, principally bedding. It also produces closely

spaced fractures and m a y completely disaggregate the swelling clay minerals. The

W o m b a r r a Shale and thefissileclaystone interbedded in the Coalcliff Sandstone, even

in fresh samples from drill holes, are particularly prone to this type of attack.
174

The potential for rock breakdown during slake-durability testing depends upon the

permeability and porosity, which control the entry and retention of water and its

mobility inside the rock; the presence of water, which can cause solution of cement,

disruption of interparticle bonds, and reduce shear strength due to increase of pore-

pressure, has a major effect on slope failure.

The solution of siderite and calcite cement in both the sandstone and claystone units

along the coastline is a contributing factor affecting breakdown of the strata. This

solution is usually a slow change, much slower than the associated physical processes;

however, over long periods of time it is more important. Breakdown of the sandstone

and some siltstone beds is primarily controlled by geological structures which allow the

entry of water. An increase in water content will lead to an increase in slaking and a

decrease in strength, which, in turn, will increase the possibility of landsliding along th

coastline.

The Coal Cliff and Scarborough Sandstone units consist of cross-bedded sandstone with

interbeds of shale and sandy shale. Marine abrasion and salt riving attacks this

sandstone just above mean sea level where it occurs at the base of the cliff and tends

to wash the clay from between the quartz grains in the sandstone beds. This causes a

significant decrease in the strength of the surface layers. Small rock fragments or

individual crystals can then fall off, or be washed off, by waves. The increased

porosity in the surface layer also facilitates the deposition of salt from evaporating sea

water. Such salt deposits expand during crystallisation and can force grains or thin

layers off the surface by repeated solution and precipitation mainly just above high tide

level.

Sandstone units along the coastline, which contain abundant clay minerals and volcanic

rock fragments, show significant strength loss on wetting. In the clay-rich varieties the

change in strength is likely to be related to the softening and expansion of smectitic

clay minerals. The petrological study in this research showed that the Narrabeen Group
175

sandstone contains common swelling clays. Therefore, expansive forces within these

rocks represent an important mechanism contributing to strength loss. High proportions

of expansive clay minerals were detected in both the volcanic rock fragments (which

were also partly replaced by chert) and in the matrix, which suggest that clay softening

in the presence of water is very important in controlling moisture-related strength

reduction in the sandstone in the niawarra area. The nature of the cement also

influences slake durability. Sandstone units in the upper Narrabeen Group, containing

mainly kaolinite and calcite cements, are generally less susceptible to moisture effects.

Therefore, slake durability of intact samples from the Narrabeen Group rocks is

controlled primarily by the mineralogy of rocks and to a lesser extent by the rock

microfractures.

The cliffs along the northern Illawarra coastline are undergoing differential erosion.

Differential erosion is the result of weathering of rocks which are not uniform in

character, but are softer or more readily altered in some places than in others. The

result is usually an uneven surface, with the softer rocks being removed more quickly

than the harder strata. Weathered sandstone in the Narrabeen Group is moderately

strong to strong, whereas measured slake durability values from weathered shale and

claystone show that these rocks range in strength from moderately weak to moderately

strong. The highly weathered shale and claystone samples were too weak to test and

became totally disaggregated on wetting.

Rock strength is related to the amount and type of weathering. In the Illawarra samples

there was a marked decrease in strength and durability from fresh to weathered

specimens. Among the weathered specimens, moderately and highly weathered Stanwell

Park Claystone samples are moderately weak to moderately strong but generally show

a very low durability. Where mudrocks are excavated and left exposed to the weather,

they rapidly slake to produce a soil-like material having significantly inferior engineering
176

properties to those of the original rocks. In these situations, the durability of mudrocks

becomes the most important factor affecting the slope stability.

The claystone interbeds in the Bulgo Sandstone and the highly weathered Stanwell Park

Claystone samples have a very low durability. This has a significant effect on slope

stability in the Bulgo Sandstone and in the Stanwell Park Claystone which is the

bedrock below m u c h of the talus mantle between Clifton and Stanwell Park.

The effect of chemical alteration (either of detrital mineralogy or matrix material) on

the effective stress and shear strength of talus soils is small in the short-term, to the

point of being inconsequential, when considering the initiation of land-sliding. The

strength of talus soils changes very little due to chemical weathering effects over h u m a n

life spans, and landsliding of these materials is primarily initiated by changes of water

pressure. A s the degree of chemical weathering increases, the proportion of fine

particles in the talus materials increases which, in turn, increases the total surface area

of the particles. With an increase in the proportion of fine particles, the void ratio

decreases, drainage potential drops and consequently water pore pressure increases.

Therefore, over the long-term, this latter factor has a very important influence on

potential slope stability. However, it should be emphasised, that these changes in

permeability and drainage potential are gradual and not important on a h u m a n time-

scale.
177
CHAPTER 8
SLOPE STABILITY IN THE NORTHERN ILLAWARRA

8.1 INTRODUCTION
In the Illawarra area, with its steep coastal escarpment, slope stability is an old problem

and can have disastrous effects on development. This is directly related to the geology

and geological history of the area. The 300 m high escarpment consists of flat-lying

Permo-Triassic volcaniclastic coal measures plus fluviatile sequences capped by a well

cemented quartz sandstone. The interbedded strong sandstone and weak shale succession

(Fig. 1.5) in the lower part of escarpment has been acted upon by erosion, stress relief,

weathering, creep and sliding processes to produce masses of talus on many of the steep

hillsides (Chowdhury, 1976). Rockslides are rare but rockfalls and toppling are common

on the steep rock slopes in the area.

The majority of slope stability problems of economic significance in northern Illawarra

involve translational or rotational slides, or slow to rapid flows of soil, talus or fill.

The precarious equilibrium of talus masses is frequently upset by heavy precipitation and

by man's activities, e.g. removal of toe support, loading the slope, and changing the

surface and subsurface drainage.

Artificial slides, many of which include underlying or adjacent talus, almost invariably

result from poor site selection or poor design and construction practices. Slope stability

evaluation in the northern niawarra should be an interdisciplinary geotechnical endeavour

requiring concepts from engineering geology, soil mechanics and rock mechanics. Of

these three disciplines, engineering geology is probably the most important. Reliable

evaluations of slope stability must begin with an understanding of regional and site

geology and of the geologic processes which formed the site and continue to act upon

it. Once this level of geologic understanding is reached, slope behaviour can often be

assessed on the basis of professional judgement, experience and precedent. Where


178

detailed stability analysis is required, the above-mentioned geologic understanding

mandatory for development of appropriate geotechnical models. This study is structur

according to the above philosophy, with emphasis on the talus slope deposits.

8.2 SLOPE DEVELOPMENT PROCESS

Various aspects of slope development in the northern Illawarra have been described b

Hanlon (1952, 1953, 1958), Bowman (1972, 1974), Amaral (1975), Chowdhury (1976),

Evans (1978, 1981), Young (1977, 1978), Walker et al. (1987), Hutton et al. (1990) and

Ghobadi (1993, 1994). Slope development processes can be simplified and generalised

as follows: stream erosion has carved longitudinal and transverse escarpment profile

reflecting local stratigraphy, and has also removed lateral and vertical support fro

escarpment walls. Stress relief accompanying lateral support removal produced tensio

fractures and bedding plane shear zones in rocks adjacent to the escarpment walls.

These stress relief features, along with stratigraphic and lithologic details, contro

groundwater flow in the vicinity of the slopes; perched water tables and hillside sp

are common. Stress relief features and related groundwater phenomena have accelerate

physical and chemical weathering of rocks on the slopes. Under these conditions,

rockfalls and topples are common. They occur on the natural slopes, with weathering

and erosion undercutting joint-bounded rock blocks which slump backward or topple

forward depending on the their geometry, support conditions, and applied forces which

in addition to gravity, often include water. Rockfall volumes are typically small,

ranging from approximately 0.1 to 24 m3. Fine-grained and argillaceous rocks

predominate in the typical stratigraphic section, so the weathering products are usu

silty clay or clayey silt with rock fragments ranging in size from sand to very larg

boulders. As weathering progresses, the strength of the near-surface soil and rock

materials decreases and they begin to creep or slide down the relatively steep

escarpment walls under the action of gravity and water forces. Deeper seated
179

landsliding along bedding plane shear zones resulting from escarpment stress relief have

also occurred in certain locations.

Eventually, these processes produced mantles of talus soil and rock fragments on m a n y

slopes. The composition, thickness and inclination of talus on a given slope reflect the

stratigraphy and erosional history of the slope. A great variety of talus slope

configurations are present. T w o cross-sections of talus slopes are shown in Figure 8.1.

W h e r e slopes are relativelyflat,on ridge tops or large erosional benches, residual soils

have formed. Residual soil on benches below steeper slope segments are often covered

with talus.

W h e r e erosion was intense, little or no talus has accumulated and rock strata are

exposed on steep slopes. Talus thickness can range from about 1 m to more than 10 m

but is typically in the order of 5-10 m in the study area. A m a x i m u m thickness

approaching 20 m occurs at the toes of slopes in thick sequences of weak rocks (e.g.

claystone and shale) where deep-seated landsliding has occurred or where accumulated

talus has not been removed by erosion. Such conditions are rare.

The most important engineering implication of talus slope development is the presence

in talus of surfaces, or zones, along which shear strength has been reduced to, and

maintained at, residual or near residual levels by a combination of softening and strain

effects. Talus soils in the northern niawarra are generally cohesive and in most cases

are fissured. Such soils containing a significant clay fraction have pronounced

tendencies for softening on wetting (Terzaghi, 1936), as well as strain-softening

(Skempton, 1964).

M o v e m e n t due to creep, sliding, or both, during slope development is generally

concentrated along one or more surfaces or zones, commonly at the soil-rock contact.

Additional surfaces or zones of movement m a y exist at levels within the talus,

particularly where talus is thick.


180

A talus derived from claystone is often finer grained than the original rock, having a

larger clay fraction and plasticity index. This is often found at the base of talus

deposits which commonly mantle the coastal terraces. As a result, in a landslide

situation, the operative strengths in the clays, which are derived from the basal clays

strata, are the residual shear strengths.

8.3 RELATTVE IMPORTANCE OF OTHER GEOLOGICAL FACTORS FOR

SLOPE FAILURES ALONG THE ILLAWARRA COASTLINE

(1) Platform recession

The study area experiences a low tidal regime (1.2-1.6 m), moderate swell conditions

and infrequent storms. Platforms are susceptible to weathering which is related to the

degree of exposure to wave assault. Minor erosion takes place along the outer edge of

the platform due to hydraulic action of waves, acting along joints, especially during

storms. Platform surfaces appear to be reasonably stable since the low permeability of

the rocks prevents water loss during low tides and the saturation enhances the stabilit

of the diagenetic mineral cements and detrital grains. However, abrasion does occur,

especially along joint surfaces, due to movement of material across the top of the

platform by wave action. Most erosion is concentrated along the base of the headland

at the back of the platform.

(2) Cliff erosion

Recession of the sea cliffs in the northern niawarra region is caused by active

weathering and wave erosion in the splash zone within and at the base of Coal Cliff

Sandstone and within the Illawarra Coal Measures at Clifton (Figs 3.22, 8.34, 8.36 and

8.40b) and within the Scarborough Sandstone in the Stanwell Park area (Fig. 8.48).

This erosion causes undercutting and subsequent collapse of vertical joint blocks.

Although the rate of cliff recession is low to the east of the slope failures, it has c
181

oversteeping of the talus adjacent to the cliff and has initiated slumping and sliding of

the talus layer progressively back from the cliff edge. This would, in turn, have a

destabilising effect on the whole of the talus layer.

(3) Erosion of headlands

According to field observations during this study, marine erosion of the cliff sequences

is concentrated at the toe of exposed headlands and is less significant in the bay areas.

Since most of the slip areas along the coastline in the northern niawarra region occur

on headland areas, this faster rate of cliff erosion is significant in initiating slips and

maintaining unstable slopes. The higher wave energy in these areas also serves to

remove fallen material from the base of the cliffs, thus preventing the build-up of a

protective toe layer of talus.

(4) Coal mine subsidence

Coal mine subsidence would act as a triggering and contributing factor to land instability

in the northern Illawarra region in that it reduces normal forces across horizontal

bedding planes and rock defects, encouraging bedding plane slip and, hence, facilitating

progressive expansion of the rock mass in an easterly direction (Fig. 3.22a, see adit).

Examination of the Coalcliff Colliery mine plan revealed that extensive coal extraction

has been undertaken in the Bulli Coal below and to the west of the Coalcliff Slip

(Fig. 6.11a). The goaf resulting from this mining would have been deflected towards

the exposed (unconfined) portion of the escarpment causing additional opening of joints

throughout the Narrabeen Group which overlies the extracted coal. This factor,

combined with the high horizontal stress regime in the region, is probably a major

contributor to the destabilisation of the rock mass along the escarpment.


182

(5) Dykes

Dykes are easily recognised at the surface, because the physical characteristics of the

dykes are different from those of the host rock. Rapid surface weathering and alteration

of the mafic dykes magnifies the lithological differences.

The dykes have commonly intruded along pre-existing zones of weakness such as faults

or more closely jointed rock masses. D y k e intrusion has locally caused hydraulic

fracturing and brecciation of the adjacent rock mass during intrusion. Thus most dykes

were intruded parallel to the major joint sets in areas where the joints are closely

spaced.

Dykes m a y affect slope stability in the study area since they: (a) change the rock mass

characteristics as they rapidly weather to considerable depths; (b) cause water ingress

into the rock mass; and (c) are generally associated with more closely jointed rock

masses.

F e w dykes are visible on the ground in the study area (see section 3.4) but some are

k n o w n from adjacent mine workings. The presence of dykes is a contributing factor

for slope instability since they rapidly weather to clay and allow water ingress at the

surface, but their effects needs to be determined with more exploratory and research

work.

8.4 ENGINEERING GEOLOGIC FAILURE MODELS FOR SLOPE


INSTABILITY ALONG THE NORTHERN ILLAWARRA COASTLINE
In the study area, especially between Scarborough and Coalcliff, crests and main slump

scarps are parallel to, and probably coincide with, vertical open fractures in the

underlying bedrock (Fig. 5.16). Along the coastline, the bedrock fractures can be seen

to have controlled the location and orientation of most small slumps. A slip surface can

develop at least partly along bedding planes in the bedrock (Fig. 8.2a), or the slip

surface m a y be confined along the contact of talus with the bedrock (Fig. 8.2b). In
183

two cases discussed in this thesis (Clifton Earth Slump, Fig. 8.34, and Harbour Slump,

Fig. 8.46), opening or movement along the 015 - 020 striking vertical fractures in the

bedrock stretched the overlying talus, forming cracks which allow ingress of surface

water and mark the crests of the resultant slides. Although the rock sequence in the

Illawarra escarpment is under the influence of a high horizontal stress (Stone, 1990;

Walton et al, 1990), the faults and fractures are not tectonically active, and no recorded

earthquakes have been reported to have epicentres related to any faults in the study area.

Hence, any recent movement along these fractures must be gravitational.

It is possible that at least some segments of the rupture surfaces in the niawarra region

m a y pass through the upper part of the bedrock, which is generally weathered and has

a relatively lower strength than intact bedrock. This type of failure was recorded in part

of the Moronga Park slump (Fig. 8.36). O n the edge of the escarpment, slight lateral

spreading of laterally unconfined blocks of bedrock, along lubricated bedding planes, can

widen bedrock fractures at the backs of the blocks, which in turn stretches the overlying

soil and forms tension cracks at the surface (Fig. 8.2a, b). Downhill movement of

blocks of rock is caused by forces exerted by expanding clays, which fill these open

fractures, as well as a sudden rise of cleft water pressure after heavy rains.

More frequently, the slip surface is the contact between bedrock and talus. The step-

like topography of the bedrock, with steps dropping at vertical fractures, forms wedges

of talus soil (Fig. 8.2b). The base of a wedge has a general downslope dip and the

back of it is marked by a fracture wall. In wet seasons, water flows along the contact

between talus and bedrock, drops at the joint steps and washes away finer grains. After

heavy rain, the water table rises sharply, reducing the shear strength along potential slip

surfaces within the talus material. The weight of saturated soil exerts a downhill

component to any expansion movement caused by swelling clays. This slight downhill

movement of soil can develop a tabular vertical gap at the back of the talus wedge,

which, in turn, stretches the overlying soil. The resulting crack marks the crest of the
184

landslip involving curved slip surfaces within the talus (Figs 8.2c, 8.39). Opening of

cracks to bedrock promotes further ingress of water and further enhances slip movement.

Most slips occur after heavy rain but they usually follow periods of wetter-than-norrnal

conditions that have already fully saturated both the rock and talus.

8.5 TYPES OF SLOPE INSTABILITY


8.5.1 ROCKFALLS AND TOPPLING
Rockfall and toppling along the escarpment usually occurs when joint-bounded blocks

of sandstone fall as a result of undercutting. They are widely-spaced along the

escarpment and do not happen frequently. Rockfalls are clearly visible along the

escarpment west of Coledale and especially between Clifton and Coalcliff.

In the northern niawarra, across the Bulgo Sandstone outcrop, debris from these

rockfalls spills down the steep upper slope of the escarpment. The large broken blocks

remain on the upper slopes until further movement occurs above. Erosion by rockfalls

along the escarpment is not a very active process under natural conditions. An

estimated rate of retreat of the cliff-line for the period 1950-1975, of 0.15 m/1000 year

may have little relevance to the long-term average (Chowdhury and Young, 1987). Cliff

retreat since the initiation of Tasman Sea rifting 100 Ma ago is estimated to be 25-

40 km with an average rate of 0.25-0.4 m/1000 years. It was probably faster in the

past when continental shelf and coastal plan were narrow and, therefore, the 0.15 mm/

years would be a very realistic rate for present retreat.

The present-day rockfalls produce limited deposits of coarse debris because the fine

matrix is rapidly washed out of the surficial deposit. They do not generate extensive

mixed deposits like the relict talus.


185
8.5.2 SHALLOW DEBRIS SLIDES

Shallow debris slides are common especially after prolonged heavy rainfalls on the

upper slopes of the escarpment. These debris slides may involve only the shallow soil

cover but also may remove weathered bedrock. They appear to be more likely to occur

after heavy rainfalls of short duration than after light rain over long periods.

These movements are typically long and narrow with planar slip surfaces. Because they

are restricted to the forested upper slopes they rarely present a problem to the inhabited

part of the area.

8.5.3 DEBRIS FLOWS

Debris flows are more common where large volumes of surface water are present. The

best known of this type failure is the State Rail Coledale Rawson Street failure which

occurred in April 1988. Failure occurred in the form of a mudflow.

8.5.4 DEEP-SEATED SLUMP-EARTH FLOWS

Deep-seated slump-earth flows off the faces of benches and the lower ridges of the

escarpment are the most extensive and severe natural failures in the northern Illawarra

area. They occur in a variety of slope materials including talus mantle, Narrabeen

Group and niawarra Coal Measures. The failure planes of these movements may be as

much as 15 m below the ground surface and they occur in materials of low shear

strength.

Upper slopes and benches of the escarpment allow water to seep deeply into the

talus/weathered rock rather than flowing rapidly off. Major slumps develop on the face

of the Wombarra Shale where this seepage emerges.

These failures are triggered by pore water pressures which develop as water moves at

depth through the slopes. They are not associated with isolated intense heavy rainfalls

where most of the rainfall flows off rapidly as overland flow. Rather they are
186

associated with periods of prolonged moderate to heavy rain and begin a few d

particularly heavy falls within such periods. Because of the strong influence

subsurface seepage these failures occur on slopes of gentle gradient. The ang

instability of slopes is between 15 and 25 in the Illawarra Coal Measures a

Narrabeen Group and between 10 and 15 in talus. Slopes gentler than these va

may alsp be unstable if groundwater conditions are unfavourable.

8.5.5 CREEP

Creep is common along the slopes of the Illawarra escarpment and is likely to

more of an issue in the future. Creep movement is difficult to detect from wa

inspections, especially if fall or regional movement has occurred. Survey tec

movement detectors such as inclinometers are often necessary to identify cree

(Huang et al, 1994).

8.6 FAILURE OF TALUS SLOPES

8.6.1 INTRODUCTION

Talus is geological material which has moved downslope under the influence of

i.e., landslide or creep debris. Talus slopes are natural slopes which have a

history of landslides or creep, or both. The zone of talus along these slopes

potentially unstable because the shearing displacements associated with past

have reduced the shear strength along the surface (or surfaces) of sliding or

both. When an excavation is made at the toe of a talus slope, failure is frequ

initiated along the existing surface (or surfaces) of sliding in the slope.

Many landslides have occurred in the northern Illawarra area, some of them ar

the Lawrence Hargrave Drive. A section of the road between Clifton and Coalcl

passes through the Wombarra Shale and a zone of talus in the wall of the esca
187

When construction began last century on this section of road, several slides were

initiated along ancient landslide surfaces in the talus. Modern landslides were

investigated along this section of road during this study. The location and geology of

the slide sites are described herein, along with engineering and geological features of

the slides.

8.6.2 ORIGIN OF THE TALUS


The origin of the talus is complicated because no reliable estimate of its age is

available. While the talus has often been considered to be of Quaternary age, the

Illawarra escarpment is no younger than Miocene and the talus may in some places also

be of Tertiary age. Duricrusted surfaces and subaerial basalts on the coastal plain below

the escarpment demonstrate that uplift of strata in the southern Sydney Basin and

formation of at least part of the coastal escarpment took place prior to 30 Ma B.P.

(Wellman and McDougall, 1974). Young commented that linear dissection has

dominated slope processes since uplift and that scarp retreat has been extremely slow.

Nevertheless, even very slow rates of retreat since the Miocene could account for the

2-3 km distance now separating the upper cliff-line of the Illawarra escarpment and

outliers of talus on the coastal plain.

8.6.3 PARENT MATERIAL OF TALUS


The talus in the northern niawarra has been derived mainly from the strata outcropping

on the middle and upper slopes of the escarpment (the Triassic Hawkesbury Sandstone

and the Late Permian - Triassic Narrabeen Group). The early Late Permian Illawarra

Coal Measures have been of minor importance as a source of talus in the northern

Illawarra. In this area talus often lies above the niawarra Coal Measures outcrop.

Therefore, the major source rocks have been claystone and quartz-lithic sandstone unit

in the Narrabeen Group.


188

The jointed Hawkesbury Sandstone has supplied most of the boulders found in the talus,

even in deposits now several kilometres distant from the cliffs. Narrabeen Group

sandstone floaters occur less frequently and the more friable and easily weathered

sandstone of the Illawarra Coal Measures are not represented in the boulder fraction.

The Hawkesbury Sandstone is highly quartzose and is very resistant to weathering.

Therefore, its boulders remain hard and angular with no sign of spheroidal weathering

or disintegration since deposition.

In contrast the sandstones of the Narrabeen Group and Illawarra Coal Measures are

more easily weathered and more likely to have disintegrated during transport. They

have a higher percentage of rock fragments; and the rock fragments are often volcanic

rather than siliceous. The most common clay mineral in the Narrabeen Group is

kaolinite with illite and mixed-layer clays also present in the lutites. Because of

weathering changes and the range of clay minerals in the parent strata, only the boulder

fraction provides direct evidence of the source of the talus. Therefore, the origin of th

unstable boulder material must result from small shallow debris flows or avalanches

from the slope above them. Falls of weathered rock, like the fall from the Narrabeen

Group mentioned earlier (section 8.3.1), produce slides with some iron staining and with

sandstone fragments in a clayey matrix.

8.6.4 CLAY MINERAL ANALYSES


Clay minerals have a significant influence on the engineering behaviour of talus. The

engineering behaviour, and especially the residual strength of talus derived from

argillaceous rocks in the northern Illawarra, depends on mineralogy. In particular, pore

fluid composition and types of cations adsorbed on mineral particles exert a significant

influence on the residual strength and on time-dependent strength changes, if any, along

existing failure surfaces in talus (Kenney, 1967; Mitchell, 1976).


189

T h e X-ray diffraction study of the talus materials, Wombarra Shale, Stanwell Park

Claystone, and interbedded clay and highly weathered sandstones in the lower Narrabeen

Group indicate that they contain quartz, iron oxides, kaolinite and expandable lattice

mixed-layer clay minerals (see chapter 4). The latter consists of randomly interstratified

illite and smectite (montmorillonite). N o significant feldspar was detected in the

previous petrological study (see chapter 4).

For more information fifteen clay samples from the five landslides were analysed. The

X-ray diffraction traces of clay samples with particles less than 2 microns show no

chlorite (Fig. 8.3a) in the matrix of the talus. Kaolinite and mixed-layer smectite clay

are the dominate clay minerals in all the talus samples (Fig. 8.3a). The limited data

suggest that expandable lattice clay minerals m a y occur preferentially along the failure

surfaces of ancient landslides in this area.

Groundwater flow through the relatively permeable shear zones m a y have caused

geochemical changes that resulted in the formation or concentration of these expandable

clay minerals, or both, along the failure surface. Further investigations are needed to

clarify this but it is noted that similar phenomena have been inferred for a failure in

colluvium derived from Carboniferous claystone at Walton's W o o d , England (Early and

Skempton, 1972).

8.6.5 GEOTECHNICAL PROPERTIES OF TALUS


Index properties were determined for samples of talus materials obtained from 15

locations (five landslides) in the northern Illawarra study area between Clifton and

Stanwell Park. They include seventy five direct shear tests, sixty liquid limit tests,

thirty plastic limit tests, fifteen determinations of in-place soil density, water-content

determination and particle-size analysis.


190

8.6.6 TEST RESULTS

The shear strength of talus has been determined from laboratory tests. The result of

tests are shown in Figures 8.4 to 8.22 and in Table 8.1.

The peak shear strength values in this study refer to samples from slipped talus

materials, moderately weathered Wombarra Shale and highly weathered sandstone. The

talus is already in a weakened or softened condition after failure. Therefore, these

measured peak shear strength values from extracted samples may be lower than the true

undisturbed peak shear strength of talus before the process of failure began.

8.6.6.1 Strength Characteristics of the Talus Matrix

Shear strength depends on physico-chemical cohesion and on interlocking and friction

between grains in clay soils (Trotter, 1993). It is affected by the pressure of pore water

filling the voids. Therefore, soils of similar texture, clay mineralogy and permeability

can be expected to have similar engineering behaviour and are generally classified by

a chart relating their plasticity index to their liquid limit (Terzaghi and Peck, 1967).

Liquid and plastic limit tests in this study showed that, based on soil classification, th

talus matrix in the northern Illawarra (between Clifton and Stanwell Park) is inorganic

clay of low plasticity (Fig. 8.23).

General relationships between the index properties of a soil and its strength have been

widely accepted. In considering the long-term stability of clayey soils, Skempton's

(1964) finding that cohesion is very small and that the angle of shearing resistance (<(>)

falls with increasing strain to a value which is constant (residual angle (|)r) has been

frequently applied. <t>r declines with increasing clay content. This decline is echoed by

a fall in (J)r with increasing plasticity index, from which the angle of shearing resistan

is usually predicted. A relationship between <j)r and plasticity index (Ip) was presented

by Voight (1973; see Fig. 8.24), Kanji (1974; see Fig. 8.25), and Lupini et al, (1981;
191

see Fig. 8.26). Also a relationship between residual friction angle with clay fraction

was presented by Lupini et al. (1981; see Fig. 8.27).

The following relationship was preposed by Kanji (1974):

((>= 46.6/L0466

The residual angles of internal friction estimated for the talus are shown in Table 8.2.

If w e assume that long-term stability of the talus is governed by the residual strength

and by seepage parallel to the ground surface, with the top flow line at the ground

surface on long straight slopes (Carson, 1969), then w e can estimate the approximate

inclination of a stable ground surface slope, p, as:

tan f3 = 1/2 tan ,

Taking (f>r to be 19, 22, 17, 15.5 and 17 (from Table 8.2), this equation yield

estimates of |3 as 9.5, 11, 8.5, 825' and 8.5 for the Clifton Hotel slump, Moronga

slump-earth flow, Jetty rock slump, Harbour slump and Coalcliff slump respectively.

These estimates are not presented here as definitive angles for the gradients of long-

term stable slopes on the talus; clearly the assumptions on which they are based m a y

not always be satisfied. Rather they are presented to indicate that the stable gradients

on the talus are low, comparable with angles suggested for clayey slopes elsewhere.

For example, Carson and Kirkby (1972) suggested 8-ll for clay slopes in general, and

Chandler (1974) noted 4 for weathered fissured Lias Clay.

Along the Lawrence Hargrave Drive between Clifton and Stanwell Park, slopes in talus

are unstable at angles above 10 and in some cases m a y be unstable at even lower

gradients depending on groundwater conditions.

8.6.6.2 Correlation of Engineering Indices and Properties

1. Atterberg limits

The relationship between the liquid limit and plastic limit is shown in Fig. 8.28). There

is a clear trend of increasing plastic limit with increasing liquid limit. Figure 8.29
192

shows a plot of liquid limit against natural water content. A rough trend can be seen

showing the liquid limit to increase with increasing natural water content.

2. Natural water content

It has been well established that the moisture content in talus can have a significant

effect on its engineering behaviour. A plot of bulk density versus moisture content (Fig

8.30; Table 8.1) shows a definite tendency for bulk density to increase with increasing

water content as the pores become filled.

3. Clay fraction

The results of grain size analyses indicate that the clay fraction, that is material fin

than 2 microns, varies between 1.60% and 26.06% (Table 8.1). For the niawarra talus,

there is a relationship between the percentage of mixed-layer (expansive) clay and the

liquid limit for the samples listed in Table 8.1. An increased clay fraction (less than

2 microns) is associated with an increase in liquid limit and an increase in plasticity

index. The poor correlation may result from many factors, for example small sample

size, differing sizes of clay particles (Hawkins and McDonald, 1992), or partial

adsorption of iron or aluminium hydroxide cations by the expanded clays (Grice et al,

1982). It emphasises the difficulty of estimating one soil property from a different

property.

Figure 8.31 shows the relationship between the angle of shearing resistance (peak and

residual) for the matrices of the niawarra talus versus the clay fraction. Figure 8.32

show the relationship between the peak angle of shearing resistance of the niawarra

talus versus percentage of clay and silt. Most of the points indicate a definite tendenc

for <|>r and <|>p to decrease with increasing content of clay particles. Figure 8.33 show

the relationship between clay fraction and liquid limit (WL%) for Illawarra talus.

Increasing clay fraction is associated with an increase in liquid limit.


193

8.6.7 CONCLUSION

The clay mineralogy can be used to identify the likelihood of instability in talus in the

northern Illawarra. The main clay minerals in the talus, claystone and shale of

Narrabeen Group are kaolinite, illite, montmorillonite and mixed-layer clays. In general,

samples with a high percentage of kaolinite are more stable than those high in illite.

Montmorillonite, although less common, was found to be less stable again in the

Illawarra area. In the detailed study area, samples with high percentages of kaolinite

are also unstable. This is related to discontinuities in the talus and existence of mixed-

layer clay minerals in the talus.

Discontinuities cause an increase in bulk permeability, allowing more rapid percolation

of water to depth. Where discontinuities penetrate to a sufficient depth, the associated

loss in shear strength may be enough to initiate slope failure and to reduce the shear

strength to its residual value (Trotter, 1993).

The geotechnical properties of talus derived from the Narrabeen Group rocks in the

northern niawarra vary widely but can be generalised. Talus ranges from GM and GC

through SM and SC to CL and ML soils in the Unified Soil Classification System

(Terzaghi and Peck, 1967). Failure surface clay usually consists of CL or ML soils and

occasionally CH, SC or SM soils.

Index properties and Unified Soil Classes give insight into the engineering behaviour of

talus. But engineering behaviour is governed largely by pre-existing discontinuities, such

as shear zones and failure surfaces. Index properties of shear zone and failure surface

materials usually differ little from those of the overlying talus, with one important

exception. Due to seepage along the relatively pervious shear zones and perhaps due

to related mineralogical and geochemical changes, many shear zones and failure surfaces

have water contents a few percent higher than those of overlying and underlying

materials.
194

The nature and geological setting of talus are such that its undrained or total stress shear

strength is virtually meaningless, except perhaps as an index property (Morgenstern,

1967). Drained or effective stress shear strength are much more meaningful in

situations involving either short term or long term stability, despite the inevitable

uncertainties associated with estimating in situ pore water pressures.

Peak and residual strengths of talus and its discontinuities depend on four inter-related

factors: (a) clay fraction; (b) amount of particles of sand size and larger (mainly rock

fragments of various types); (c) soil structure which includes fabric, composition and

inter-particle forces (Mitchell, 1976); and (d) degree of weathering or alteration.

Augmentation of talus by fine particles increases the total surface area of the particles

and consequently increases the angle of internal friction. With an increase in the

proportion of fine particle, void ratios decrease. Drainage potential drops and the

likelihood of developing positive pore water pressure consequently increase. Despite the

increase in the angle of internal friction of this material, the changes to hydrological

properties have an overriding influence on potential stability. In first time slides, the

value of the shear strength is close to peak shear strength, whereas in second or later

slides along the same slip surface the value of the shear strength is close to the residual

shear strength (Skempton, 1964). Residual angle (tyt) values for talus matrix are

between 15.5 and 22.

Calculation of shear strength parameters required for the limiting equilibrium of talus

masses is subject to significant uncertainties regarding in situ pore water pressures and,

to a lesser extent, slide mass geometry. Nevertheless, if the geological framework is

well understood, it is usually possible to calculate sets of strength parameters

corresponding to the limiting equilibrium of talus masses for reasonable bounding

conditions of water pressure and geometry. Interpretation of calculated strength

parameters, within the geological framework of the slope and known range of talus
195

strengths, often permit the range of possible strength value to be narrowed sufficiently

for use, with judgement, in design.

The low ((>, values for talus deposits in the northern niawarra may give some confidence

in the above techniques for evaluating stability of talus slopes but they also point out

the need for field observations and monitoring of water pressures in these slopes. Talus

failures that were observed on talus slopes in the northern Illawarra appear, on the basis

of literature review, to resemble similar features in colluvium derived from the

argillaceous rocks in England, such as at the famous Walton's Wood landslide (Early

and Skempton, 1972) and colluvium derived from the argillaceous rocks in Western

Pennsylvania (Hamel, 1969, 1980).

8.7 CASE STUDIES

The study area site plan (Fig. 8.34), shows three regions containing seven study sites.

In general, distinction between bedrock and intra-talus surface failures needs subsurface

information (drilling or geophysical investigation) but it is certainly possible distinguish

common bedrock blocks in the talus mantle along the coastline (e.g. Fig. 3.10).

Bedrock blocks are from the upper niawarra Coal Measures in the south and from the

lower Narrabeen Group in the north, respectively. For example in the Clifton earth

slump bedrock is the Coal Cliff Sandstone and the failure surface is in the talus and

weathered Wombarra Shale (Figs 8.35, 8.36). In the Jetty rock slump the failure surface

is related to the Jetty Fault (a series of small en echelon faults in the Jetty Fault

suggests that this fault may have been active during deposition, Fig. 2.11). In the Jetty

rock slump part of the surface of the Jetty Fault acts as a failure surface in the Coal

Cliff Sandstone (Figs 8.45, 8.46).


196
REGION A

8.7.1 SITE 1 CLIFTON EARTH SLUMP

8.7.1.1 Location

This slump is located beside the Clifton Hotel. The area is on a small flat portion of

ground on the east side of the Lawrence Hargrave Drive above the sea cliffs (Fig. 8.35).

8.7.1.2 Geology

The town of Clifton is located on a narrow coastal terrace within the steep slope of th

escarpment above the coastal cliff-line. The natural landforms generally slope to the

south-southeast at a gentle angle of approximately 12 to 17 both uphill and downhill

of the road respectively (Ghobadi and Pitsis, 1993).

The rocks at the site are essentially flat-lying strata consisting of alternating beds

sandstone, shale and claystone. The talus mantle comprises a heterogenous mixture of

angular, gravel to boulder size sandstone fragments with variable amounts of sand, silt

and clay. The average size of boulders is about 50 cm. Information from drilling done

by the Department of Main Roads (DMR) shows the talus to consist of yellow brown

and brown sandy clay with medium plasticity, medium to coarse sand with traces of

fine gravel, and some highly weathered sandstone fragments.

The head of the slump is underlaid by the upper beds of the Wombarra Shale (Fig.

8.36) close to the junction with the overlying Scarborough Sandstone which is visible

in a small cutting on the west side of the road. The Scarborough Sandstone contains

numerous vertical joints; this is an ideal condition for the development of contact

springs. Surface water infiltrates the Scarborough Sandstone and seeps down through

it to the contact with the relatively impermeable Wombarra Shale. This seepage

increases water pressures in the talus and contributes to alteration of the Wombarra

Shale. Both of these effects contribute to further landsliding (Fig. 8.36). The base of

the slump near the edge of the cliff lies on the Coal Cliff Sandstone and water drains
197

from the base of the colluvium just above the Wombarra Shale-Coal Cliff Sandstone

contact.

8.7.1.3 Description of the slump

The effects of movement, observed along the Lawrence Hargrave Drive beside the

Clifton Hotel, is a rotational slide (Fig. 8.35) with a tension crack at the ground surface.

The dip of the surface of sliding is between 8 and 10 toward the sea and subsidence

movement on the road at this site coincided with the April rains of 1988. It is

considered likely that the sliding surfaceflattenedor became convex-upward because this

slump moved very slowly. The slump is about 65 metres long and 40 metres wide.

The crown, main scarp, head and toe of the slump are recognisable, but the flanks are

obscured under dense vegetation. The main scarp is low and lies along Lawrence

Hargrave Drive. O n e creek occurs beside the slump to the south and has a direct

influence on slump by the introduction of water (Fig. 8.35). The main body of the

slump comprises talus and weathered Wombarra Shale. With any heavy rain, water

percolates into the slump and mobilises the matrix of the talus. O n e can expect the

formation of cracks as the slip develops further, including en-echelon, longitudinal and

crescentic cracks which extend back to the base of the Scarborough Sandstone.

Based on drilling information, between 4.5 m and 5.9 m depth, there is high plasticity

clay with highly weathered shale layers at or near the slip surface. A seepage zone is

possibly present in this area at 5.5-6 m . Bedrock is at a depth of 6 m and consists of

moderately to highly weathered jointed siltstone. This siltstone continues to a depth of

7 m . Joints in the siltstone range from horizontal to a m a x i m u m dip of 60. Spacing

between joints is between 4 c m and 10 c m and the joints are infilled with clay. The

strength of the siltstone is very low (Is(50) = 0.03-0.1 M P a ) .

Three inclinometers and two piezometers were installed at this site by the Department

of Main Roads ( D M R ) . These inclinometers show a slip plane has developed at depth
198

between 4.5 m and 5.5 m in clay, immediately above the bedrock. The piezometers

show that the groundwater table ranges between 1.5 m and 4.5 m at this site. At the

present time the slump is active and it is moving toward the sea very slowly; the

direction of movement is southeast (strike = 055, or movement direction = 145).

8.7.1.4 Geotechnical properties of the talus

Index properties were determined for samples of talus materials obtained from 3

locations (crown, head and toe) in this slump. The range and average values of these

index properties are given in Table 8.1. It must be noted that these values of shear

strength parameters do not represent the properties of soil at the slip surface. Very

careful drilling, undisturbed sampling and testing would be required to obtain the

operative shear strength parameters on the actual slip surface. No resources for these

detailed studies were available for this Ph.D. thesis.

According to the results from the tests, the peak friction angle (p) decreases from 45

in the crown to 40 in the head and to 23 in the toe of the slump (Fig. 8.7). Probably

the value in the toe of the slump represents a residual friction angle since this materi

has moved. Also residual shear strength is primarily dependent on mineral composition

and system chemistry (Kenney, 1967). Mineral composition affects atterberg index

parameters, so it does not seem surprising that strength and plasticity can be correlate

(Fig. 8.31). Mineral composition is valuable in studying the residual strength of a

particular variable soil deposit, provided that they properly reflect changes in the mor

fundamental properties of shape, grading, mineralogy (Lupini et al, 1981).

8.7.1.5 Conclusions

The slump occurred in a weak zone composed of talus, claystone and shale. The failure

surface of the slump was located at the base of the talus or just into the Wombarra

Shale. Movement of the slump reduced the shear strength along the failure surface to
199

a residual level. The strength of the failure surface materials probably remains at or

very near residual levels until times of very heavy rainfall (West, 1994).

One of the causes of the instability in this slump is erosion of the toe of the slope by

the sea, resulting in relaxation of the materials above (Fig. 8.36). The movement

detected by inclinometers is aided by the high perched water table above the Wombarra

Shale. For improving short to medium term stability in this area the water table must

be lowered to or below the top of the bedrock by drainage either out to the edge of the

cliff or downwards into the Coal Cliff Sandstone.

8.7.2 SITE 2 MORONGA PARK SLUMP-EARTH FLOW

8.7.2.1 Location

This slump-earth flow is located in the Clifton area, opposite the School Parade and east

of Moronga Park. It was initiated during the April rains of 1988 and has remained

active until now (Fig. 8.37). This slip is a retrogressive landslide but retrogression has

been slow during the last few years because of low rainfall.

Comparing an aerial photograph taken by B.H.P. in 1988 and my field observations in

October 1994 the average rate of retrogression is 25 cm per year. This rate may

increase dramatically in the future if rainfall is exceptional high.

8.7.2.2 Geology

The rocks at the site are lower Narrabeen Group and upper niawarra Coal Measures.

They are essentially flat-lying strata consisting of repeated beds of sandstone, shale, coal

and claystone (Figs 8.38 and 8.39). The head of the slump is underlaid by the Illawarra

Coal Measures.

The talus consists of silty clay, sandy clay and gravel of medium high plasticity and

mottled light grey, red and brown colours. Large sandstone boulders were encountered

in the main body of the slump (Fig. 3.10). The typical thickness of the talus may be
200

about 10 m . Underlying the talus soils are weathered coal measure rocks consisting of

tuffaceous sandstone, carbonaceous siltstone, claystone and minor coal seams. Coal

seams are found to often provide groundwater paths within the Illawarra Coal Measures

rocks.

The toe of the slump is in the upper Illawarra Coal Measures (Fig. 8.40). Water flows

over the surface of the slope and then permeates the talus, infiltrates the tuffaceous

sandstone and seeps d o w n through it to the contact with the next relatively impermeable

claystone. The water then moves laterally along this contact to the ground surface

where it emerges as a line of springs. This seepage increases water pressures in the

talus and contributes to further landsliding. The Clifton Fault is located to the north of

this slip (Fig. 8.38). This fault has an approximate east-west strike with a steep dip

(85) to the north. A tension crack regularly opens up along the strike of the fault

during prolonged periods of wet weather due to water movement along the weathered

fault plane. The past performance along this fault would suggest that it is a major zone

of weakness. Consequently, there is a possibility that any such movement could have

an influence on the slump, either directly by transference movement or indirectly by the

introduction of water (Ghobadi, 1993). This fault also creates an additional area of

instability near the Lawrence Hargrave Drive. Recent movement within Rube Hargraves

Park (Fig. 8.38) to the northwest of the slip, and the first house near the railway

corridor, is believed to be attributed to a zone of weakness association with the Clifton

Fault.

8.7.2.3 Description of the slump-earth flow

This slump is about 100 m long and 50 m wide. The crown, main scarp, head, flanks

and toe are clearly recognisable. The main scarp is about 12 m high, it is a nearly

vertical scarp surface stained by iron oxides. The toe of the slump-earth flow extends
201

onto the niawarra Coal Measures, comprising interbedded shale, coal and sandstone

(Figs 3.10, 8.40).

The natural talus slope lies on a typical talus deposit. The talus body shows clear

separation of components by grain size. The larger, heavier blocks are mostly toward

the base of the main body of the slump, which comprises soil and weathered boulders.

Wave action has winnowed some of the distal toe of the talus slope.

The morphology of these earth flows and slumps are the result of the interaction

between mass movement and the underlying geology. The flows developed in an older

talus body which came to rest above a bed of highly weathered sandstone and claystone

of the niawarra Coal Measures. The upper surface of this sandstone bed forms the

basal shear surface. The presence and continuity of low angle joints in the weathered

Illawarra Coal Measures beneath the talus confirms that the main failure plane is located

in the upper niawarra Coal Measures. Critical groundwater conditions are often a direct

consequence of the structural defects. Coal beds in the Illawarra Coal Measures act as

aquifers, with claystone units acting as aquitards. Landsliding is commonly related to

the presence of the aquifers.

The rate of movement is uncertain and likely to be erratic depending on rainfall. Based

on observations during the three years of study, the average rate of movement is 20 mm

to 30 mm per year.

8.7.2.4 Geotechnical properties of the talus

Index properties were determined for samples of talus materials obtained from 3

locations (crown, head and toe) in this slump. The range and average values of these

index properties are given in Table 8.1. As noted before these values represent peak

shear strength parameters and not the properties at soil of the slip surface.

According to the result of the tests peak friction angle ((|)p) decreases from 38 in the

crown to 33 in the head and to 30 in the toe of the slump (Fig. 8.10). The reduced
202

value for the head of the slump is dependent on the higher percentage of clay fraction

(Table 8.1) in comparison to the peak friction angle on the crown of the slump. The

value in the toe of slump represents peak friction angle in the earth flow (Fig. 8.37).

8.7.2.5 Conclusions

The slump-earth flow occurred in a weak zone of talus and clay-rich stratigraphic

section composed of the upper Illawarra Coal Measures (claystone, tufaceous sandstone

and coal seams). The main failure surface of the slump-earth flow is located in the

jointed and weathered upper Illawarra Coal Measures. The bedrock is highly fractured

and weathered. The geological and geomorphological field investigations have shown

that jointing and faulting in combination with weathering, steep topography, water

incision and heavy rainfall, played a significant role in the movement of the slump-

earth flow.

The presence of the Clifton Fault and joints in the weathered rock mass contributed to

the mass movement. Movement reduced the shear strength along the main failure

surface to a residual level, where it probably remained until times of very heavy rainfall.

Geotechnical experience with this type of landsliding generally shows that groundwater

pressure is often the dominant factor activating landsliding, when compared with soil

strength parameters and slope angle. Based on observations in the field, groundwater

pressures are present in the crown area of the landslide and these pressures are assessed

to be responsible for driving the landslide. N o laboratory testing of slide plane material

has been undertaken to date (because the depth of sliding is not accurately k n o w n at

present). Coal beds in the niawarra Coal Measures act as aquifers, and water flow

within the coal beds is concentrated along discontinuities at the base of coal unit. This

increases the rate of weathering of the already highly weathered sandstone and leads to

the next slump.


203

Due to the sensitivity of the slide to fluctuations of water table level, the remedial

works can be basically aimed at draining the slope and preventing surface water

infiltration.

REGION B

8.7.3 SITE 3 SOUTHERN AMPHITHEATRE COMPLEX LANDSLIDE

(GRABEN A)

8.7.3.1 Introduction

The southern amphitheatre complex landslide consists of large planar slides involving

both block and wedge failures, which occur within the less deformed cover strata (cf.

Pettinga, 1987a, b). They are controlled by grain size variation in the alternating

sandstone-mudstone successions, which permit the development of high pore pressures

at potential bedding-plane failure surfaces (Pettinga and Bell, 1991). The failure was

apparently triggered by this water pressure which existed in the slope. The rear of the

failure mass was defined by a vertical joint at the crest of the slope. This joint began

to open several years prior to the slide. Most of the failure surface passed through

layers of soft claystone and shale at the base of the slope.

8.7.3.2 Location

This landslide is located immediately north of the Jetty Fault along the Lawrence

Hargrave Drive between Coalcliff and Clifton (Fig. 8.41a, b, c).

8.7.3.3 Geology

The rocks consist of sandstone, with interbeds of claystone and shale overlying the Bulli

Coal.

The Bulgo Sandstone consists of grey quartz-lithic sandstone with minor reddish-brown

claystone and some thin conglomerate bands. It has an average thickness of


204

approximately 130 m in the study area. The sandstone beds are normally between

0.3 m and 3 m thick and the claystone interbeds are normally between 0.1 m and 1 m

thick (Fig. 3.9). Joint spacing averages 0.3 m. The Stanwell Park Claystone is

approximately 40 m thick (Fig. 8.42) and consists of red-brown and greenish claystone

with two lensoidal quartz sandstone beds between 0.1 m to 1 m thick. The Scarborough

Sandstone consists of thickly bedded sandstone with conglomerate beds and minor shale

(Fig. 3.6). It outcrops boldly, forms the major part of the coastal cliffs and is

approximately 25 m thick.

Wombarra Shale is about 40 m thick and consists of predominantly grey shale but it

contains one sandstone layer 6 m thick known as the Otford Sandstone Member. The

Coal CUff Sandstone is a massive grey lithic sandstone which forms the lower coastal

cliffs in the landslide area. Joints are spaced up to 10 m apart but are normally space

at 1 to 3 m. In the landslide area this sandstone is about 8 m thick.

The niawarra Coal Measures is a sequence approximately 250 m thick extending

downwards from the base of the Coal Cliff Sandstone. The uppermost layer is the Bulli

Coal which, in the landslide area, only occurs to the south of the Jetty Fault. Below

the Bulli Coal the unit consists of claystone, shale (Fig. 3.22) and lithic sandstone d

to the base of the cliff. The Bulli coal has been mined to the west of the slip area

(Fig. 3.22).

The sandstone beds are permeable because of jointing. Claystone and shale beds are

generally quite impermeable. Perched water tables are common where these

impermeable beds underlie the more permeable beds.

The Jetty Fault is located to the south of this landslide. The dip of the fault plane i

70 N on Lawrence Hargrave Drive and 45 N in the lower coastal cliff north of the old

Coalcliff adit; its strike is east-west. This fault appears to be responsible for

groundwater circulation under the road. It has a throw of 8 m at the level of the Bulli

seam which dies out to zero within the Scarborough Sandstone. This change in
205

displacement is explained by considering that the fault was only active during deposition

of the sediments. This fault has had an influence on the landslide indirectly by the

introduction of water. It also creates an additional area of instability to the east of

Lawrence Hargrave Drive (Jetty rock slump) which will be discussed later. Fault 2 (F2)

is located to the north of the landslide. The dip of the fault plane is about 45 S W , its

strike is northwest-southeast. Fault 1 (Fl) is to the south of landslide and it is parallel

to the Jetty Fault. The dip of the fault plane is about 55 N (Fig. 8.41a). The

sandstone beds (Bulgo Sandstone and Scarborough Sandstone) overlie claystone and

shale and have numerous vertical joints almost parallel to the face of the slope. These

joints are attributed to stress relief in the escarpment walls. The mechanism of this

escarpment wall stress relief and joint formation has been discussed in chapter 5.

8.7.3.4. Description of the slide

Very little slide m o v e m e n t seems to have occurred after the initial slump (Jetty rock

slump). The location of the failure surface was determined from surface indications

(Fig. 8.42). The lower part of the failure surface generally followed the base of the

W o m b a r r a Shale. The upper part of the failure surface met the open vertical joint in

the Scarborough Sandstone (Fig. 8.42). Normally, surface water infiltrating the upper

part of the slope would pass d o w n through the vertical joints in the Bulgo Sandstone

and drain laterally to the surface of the cut through the W o m b a r r a Shale. Water in the

middle to lower part of the slope would likewise drain laterally along the three fault

(Jetty Fault, Fl and F2, Fig. 8.41a). It is likely that infilling material (clay) essentially

plugged most of the natural drainage outlets in the slope face and that hydrostatic

pressures in the open joint at the rear of the failure mass trigged the slide. The long

term water seepage to the north of this landslide indicates a zone of high fracture

permeability. It confirms the existence of a penetrative rock defect, such as a fault (F2)

or shear zone beside the landslide.


206

Generally water flow within the rock mass is concentrated along discontinuities at the

contact between the Bulgo Sandstone and Stanwell Park Claystone at the middle of

southern amphitheatre and at the contact between the Scarborough Sandstone and

Wombarra Shale at the bottom of southern amphitheatre. This increases the rate of

weathering of the Wombarra Shale, causes fretting and weathering of the sandstone and

leads to toppling failures and rock falls.

8.7.4 SITE 4 NORTHERN AMPHITHEATRE COMPLEX LANDSLIDE

(GRABEN B)

8.7.4.1 Location
This landslide is located beside the Harbour Fault along the Lawrence Hargrave Drive

south of Coalcliff (Fig. 8.43).

8.7.4.2 Geology

The succession is composed generally of interbedded sandstone and claystone and is

similar to that in southern amphitheatre (Fig. 8.41a, b). The Harbour Fault forms a

prominent indentation in the coastal cliffs where it has a vertical displacement of 20 m .

It has a strike of 080 and dips 62 toward the south. This fault also appears to be

responsible for groundwater circulation under the road. It has had an influence on the

landslide indirectly by the introduction of water. This fault also creates an additional

area of instability in the east of Lawrence Hargrave Drive (Harbour Slump) which will

be discussed later. The fault 3 (F3) is to the south of this landslide. The dip of the

fault plane is about 45 N W , its strike is southwest-northeast.

8.7.4.3 Description of the slide

The whole northern amphitheatre has been considered to be moving towards the sea as

a massive landslide. The possible plane of failure would be in the W o m b a r r a Shale


207

(Fig. 8.44). A major fault to the north of the landslide (Harbour Fault) and F3 to the

south of the landslide are good localities for water flow. Movement of this type would

be expected to increase dramatically in the future. This amphitheatre area could in fact

represent or develop into a deep seated but slow moving rock slide. The slow

movement can be related to: (a) serious slippage of rock support along the amphitheatre

cliff line: (b) collapse of the underlying coal workings; and (c) the long term water

seepage at the north of this landslide which indicates a zone of high fracture

permeability.

This landslide comprises surficial mass movements, within the soft rock (Stanwell Park

Claystone), and a variety of complex failures can be recognised, including earthflow,

debris flow and talus slump. The Stanwell Park Claystone contains appreciable

montmoriUonite clay (laboratory swelling tests on the fresh claystone core samples

produced unconfined swelling strains up to 10% and swelling pressures of 100 KPa for

Stanwell Park Claystone, Railway Authority, 1983), and is prone to slaking (chapter 7)

when subject to moisture variation above the perched water. From shear strength values

compiled by Pitsis (1992), it is clear that the Stanwell Park Claystone has a generally

low friction angle. The magnitude of the friction angle may be as low as 11 (Table

8.3). Claystones are thus readily transformed into material with clay soil properties,

such as plasticity and volume change, whilst the associated siltstone and fine sandstone

readily disaggregate when saturated and disturbed.

Intense fracturing accompanying normal faulting (Harbour Fault) has additionally

weakened the rock mass, greatly assisting the slaking process by permitting rapid

variations in moisture content. Surficial mass movements occur extensively and they

involve the entire length of a slope from ridge crest to amphitheatre floor. Generally

water flow within the rock mass is concentrated along discontinuities at the contact

between the Bulgo Sandstone and Stanwell Park Claystone at the top of northern

amphitheatre. This increases the rate of weathering of the Stanwell Park Claystone,
208

causes fretting and weathering of the sandstone and leads to toppling failures and rock

falls.

8.7.4.4 Conclusion

T w o grabens (A and B ) containing two amphitheatre complex landslides are located

between Clifton and Coalcliff along the Lawrence Hargrave Drive. Amphitheatre

complex landslide associations are identified which reflect mass m o v e m e n t controlled

by bedrock lithology, rock mass structure and climate. The southern amphitheatre

complex landslide is a deep-seated block and wedge slide on the gently deformed lower

Narrabeen Group (alternating sandstone, claystone, shale) and upper niawarra Coal

Measures (Bulli Coal). The northern amphitheatre complex landslide is a deep-seated

block slide accompanying surficial creep earthflows and debris-flow slides on the lower

Narrabeen Group. Shallow talus failure, which is independent of bedrock, is normally

triggered by high intensity rainfall.

Landslides were triggered by the water which accumulated in the slopes. Water flow

within the rock mass is concentrated along discontinuities at the base of lower

Narrabeen Group sandstones.

Opening of the vertical joint at the rear of the failure mass occurred progressively over

a period of several years but the failure itself occurred suddenly w h e n water pressure

in the slope reached a critical value. These critical pressures probably corresponded to

water level of the head of the slump is at or near the top of the open joint.

The failure surface was largely defined by structural features in the slope. The open

vertical joint defined the rear of the failure mass and the lower part of the failure

surface followed two weak strata (Stanwell Park Claystone and W o m b a r r a Shale) in

the slope.
209

These amphitheatre complex landslides rotated slowly and incrementally (creep) on their

failure surfaces. These incremented movements are characteristic of failure surfaces

along which strength has been reduced to a residual value.

To produce a comprehensive engineering geological failure model for the southern and

northern amphitheatre slides, that would account for the deformation and failure

mechanisms along the large creeping rock slopes, data from surface surveys would have

to be combined with data from exploration adits and a series of carefully positioned

boreholes. Such data, to be collected from future research, would probably show that

the slopes have been deformed and relaxed. Boreholes through the slopes would obtain

rock quality designation (RQD) values for the major rock units and would permit an

assessment of the fissility in the Wombarra Shale unit (Fig. 3.15). The presence of

extensive near-surface fissility, in combination with unstable cements and swelling clays,

would account for the rapid weathering, disintegration and reduced overall rock mass

quality in this unit.

Further drill-supported field investigations should reveal that deformation of whole slopes

resulting from escarpment stress relief may have occurred in some locations.

In the upper parts of slides in both amphitheatres, tensile cracking zones, about 40 m

wide and composed of a series of vertical joints, can be observed. Within these zones

bending-toppling failure of the rock mass has occurred near the surfaces, and has

resulted in depression zones, about 10 m wide, at the backs of the slides. Tensile

cracking zones at the backs of these slides could not have resulted solely from

superficial bending-toppling deformation occurring at the top of the slopes. Rather it

appears to be associated with depth-creep deformation of the whole slope.

In summary, it can be concluded that the creep-formed slides probably have a composite

deformation mechanism, which is mainly controlled by creep along major faults which

bound these slide areas (faults 1 and 2 in the southern amphitheatre [Fig. 8.42], and

fault 3 and the Harbour Fault in the northern amphitheatre; Fig. 8.44), and creep-
210

induced tensile fracturing of the rock mass within the slide area. Slides in both

amphitheatres thus appear to fail due to progressive fracturing of the rock mass.

It is apparent that the slip failures on both slopes are still developing, so a basic

prerequisite of stability assessment of creeping slides needs to thoroughly investigate

the full course of slope deformation, that is the mechanism of deformation and failure.

Finally, using the above mechanism and present stage of slope deformation, a qualitative

assessment of slope stability can be made, and its future development and final failure

(sliding) form can be predicted.

Case studies used in this research also suggest that future research should include not

only systematic engineering geological investigations and geological surveys, but also

detailed exploration, rock mechanics tests and numerical simulation. Risk assessment

and prediction of movement in such complex slides is by no means an easy task.

8.7.5 SITE 5 JETTY ROCK SLUMP

8.7.5.1 Location

This slump is located at the toe of the southern amphitheatre complex landslide (Graben

A ) between Clifton and Coalcliff beside the Jetty Fault and east of Lawrence Hargrave

Drive above the sea (Fig. 8.45).

8.7.5.2 Geology

The rocks are essentially flat-lying strata consisting of Wombarra Shale, Coal Cliff

Sandstone and Bulli Coal. The middle beds of the Wombarra Shale are visible at the

head of the slump in a small cutting on the east side of the road. The Wombarra Shale

contains numerous vertical joints; this is an ideal condition for the development of

contact springs. Water drains from the toe of the rock slump at the contact between

the Coal Cliff Sandstone and Bulli Coal. The thickness of the Coal Cliff Sandstone and
211

Bulli Coal is about 5 m and 1.5 m respectively in this rock slump. The Jetty Fault

causes an increase in local water flow and is directly related to this rock slump.

8.7.5.3 Description of rock slump

The rock slump consists of a talus mantle, and middle and lower beds of Wombarra

Shale and Coalcliff Sandstone. The base of rock slump is in the Coalcliff Sandstone,

where it was displaced by the Jetty Fault.

The slump is about 60 m long and 50 m wide, the main scarp is steep and

approximately parallel to Lawrence Hargrave Drive. The crown, head, flanks, and toe

of the slump are recognisable.

This rock slump is a wedge slide failure and involves translational movements on low

shear strength planes that dip at angles as low as 15. Surface water infiltrates though

the talus and seeps down through it to the contact with the relatively impermeable

Wombarra Shale (Fig. 8.46). Seepage increases water pressures in the talus and

contributes to alteration of the Wombarra Shale. The failure surface can be lubricated

by montmorillonite clays, whilst water moves laterally along two discontinuities (Fig.

8.45). Both of these effects contribute to further landsliding.

8.7.5.4 Geotechnical properties of the talus

Index properties were determined for samples of talus materials obtained from 3

locations (crown, head and toe) in this slump. The range and average values of these

index properties are given in Table 8.1. It must be noted that these values of shear

strength parameters do not represent the properties of soil at the slip surface.

According to the result of tests peak friction angle decreases from 47 in the crown to

42 and 40.5 in the head of the slump (Fig. 8.13). These angles are higher than the

peak friction angles ((J)p) in the Moronga Park slump. This is relation to composition of

the talus materials. In this slump, the talus contains cemented shale fragments of
212

Wombarra Shale. The Wombarra Shale has a calcite cement in this location, which

results in the high peak friction angle found in this rock slump.

8.7.5.5 Conclusion

The main causes of slumping are, erosion of the toe of the slope by sea, water

percolation into the slump along the Jetty Fault, and mobilisation of the material after

any heavy rain in the area. This slump is an active but it is moving very slowly

toward the sea. During the last three years hardly any creep m o v e m e n t has been

observed on the basis of surface markers.

8.7.6 SITE 6 HARBOUR SLUMP

8.7.6.1 Location

This slump is located immediately south of Coalcliff, east of Lawrence Hargrave Drive

along the road and above the sea (Fig. 8.47).

8.7.6.2 Geology

The rocks consist of the Wombarra Shale and Coal Cliff Sandstone. The upper beds

of the Wombarra Shale are overlaid by the Scarborough Sandstone which is visible on

the west side of the road. Water flow within the Scarborough Sandstone is concentrated

along joints at the contact between this sandstone unit and the relatively impermeable

Wombarra Shale. This seepage increases water pressure in the fill (ashes, slag and soil)

and talus material and causes alteration of the Wombarra Shale. Both of these effects

contribute to instability of the slope.

The Harbour Fault is located immediately south of this slump. This fault has an

approximately east/west strike with dip of 70 toward the north. This fault has created

an additional area of instability to the south of this slump.


213

8.7.6.3 Description of the slump

This slump is about 25 m long and 65 m wide. The crown, main scarp, head, flanks

and toe of slump are recognisable; the main scarp is sharp and lies along the outer edge

of Lawrence Hargrave Drive. The base of the slump lies on the Coalcliff Sandstone

and water seeps from the toe of slump (between the base of the slip and the sandstone

(Fig. 8.48). The main body of the slump comprises fill and talus (weathered boulders

in a clay matrix). This slump can divided in two parts. The northern part of the

slump, comprising a shallow fill/talus failure involving the weathered mantle, occurs

irrespective of the bedrock lithology and is triggered on the steep slope (about 30) by

high intensity rainfall. The southern of part the slump is a creep mass involving talus

and weathered Wombarra Shale; it is moving very slowly southeast toward the sea.

8.7.6.4 Geotechnical properties of the talus

Index properties were determined for samples of talus materials obtained from 3

locations (crown, head and toe) in this slump. The range and average values of these

index properties are given in Table 8.1. As noted before these values of shear strength

parameters do not represent the properties of soil at the slip surface.

According to the results of tests the peak friction angle ((J)p) decreases from 44 in the

crown to 30 in the head and 16 in the toe of the slump (Fig. 8.13). This is related

to the grain size of talus materials which decreases from the crown to the toe of the

slump but not the clay content. Talus in the crown of the slump mostly consists

fragments of the Wombarra Shale as noted before for the Jetty rock slump. The value

in the head of slump represents a peak friction angle for soil fill and the value in the

toe of the slump show residual friction angle since this material has moved toward the

sea (Fig. 8.47).


214

8.7.6.5 Conclusion

The slump comprises a shallow fill/talus failure in the north, which

bedrock control, and a creep mass in the southern part of slump. It

heavy rainfall. The toe of the slump is being eroded by the sea, res

of the material above. Also with any heavy rain, water percolates in

mobilises the material. This slump is active but the southern part o

stabilised by remedial measures (retaining wall). Most of the slide a

slowly (creep) southeast toward sea.

REGION C

8.7.7 SITE 7 COALCLIFF SLUMP

8.7.7.1 Location

This slump contains a broad active slip zone and is located between

Stanwell Park between the Railway and Lawrence Hargrave Drive (Fig. 8

8.7.7.2 Geology

The sedimentary rock units exposed in the vicinity of this slump bel

Narrabeen Group. The two members of immediate relevance are the Bulg

the base of which is exposed about 4 m above the railway level, and

Claystone which underlies the railway embankment and talus slope (Fig

Stanwell Park Claystone is approximately 40 m thick at this location

three main claystone intervals and two sandstone intervals. The Bulg

overlies the Stanwell Park Claystone and is approximately 130 m thic

consists of grey quartz-lithic sandstone, medium- to coarse-grained

conglomeratic bands. A chocolate brown claystone bed about 1.5 m thi

near the base of the Bulgo Sandstone uphill of the slip area. The cl
215

is composed mainly of quartz, and kaolinite, with montmorillonite and illite. This

composition is similar to the Stanwell Park Claystone.

The dominant rock defects apparent in the bedrock exposed on the uphill side of the

slump belong to a close spaced joint set oriented subparallel to the cutting and the

hillslope. A n outcrop at this location exhibits the joints, spaced as closely as 5 m m

to 20 m m , striking between 050 to 060 with dip of between 70 and 80 to the east.

The close spaced joints were produced by weathering and stress relief subparallel to the

hillslope.

8.7.7.3 Description of the slump

This slump is about 150 m long and 600 m wide. The crown and main scarp are

recognisable. The main scarp lies along the railway line. The main body of the slump

comprises fill and talus, and base of slump lies on the Stanwell Park Claystone.

Evidence of artesian groundwater flows was discovered in February 1972 during the

diamond drilling by the State Rail Authority. From examination of drilling records it

would appear that the water flow originated from a soft sandstone layer (between 5.3 m

and 6.4 m depth) within the top of part of the Stanwell Park Claystone. The layer was

isolated above and below by grey shale beds. The core of the Stanwell Park Claystone

bedrock was highly fractured and displayed highly ferruginised joints indicating the

presence of water. This flow m a y provide significant piezometric pressures at the

failure surface and cause saturation of the colluvial mantle. In addition to this, it is

possible that a fault detected in the nearby mine workings m a y crop out in the vicinity

of the slip and result in an altered groundwater regime.

8.7.7.4 Geotechnical properties of the talus and Stanwell Park Claystone

Index properties were determined for samples of talus materials obtained from 3

locations (crown, head and toe) in this slump. The range and average values of these
216

index properties are given in Table 8.1. It must be noted that these values of shear

strength parameters do not represent the properties of soil at the slip surface.

According to the result of tests peak friction angle (<|>p) decreases from 34 in the crow

to 29 in the head and to 18 in the toe of the slump (Fig. 8.20). Talus in the crown

of the slump consists of claystone rock fragment (Stanwell Park Claystone). Depending

on the degree of breakdown of the claystone rock fragments in the talus materials the

difference of the peak friction angle occurs in this slump. The low peak friction angle

in the toe of the slump is related to the high percentage of silt and clay (Table 8.1).

The slump contains a broad active slip zone of talus, but the actual location of the

failure plane is somewhere within the uppermost weathered zone of the underlying

Stanwell Park Claystone and not in the talus layer itself. Cores recovered from diamond

drilling by the State Rail Authority (1982) showed that within this zone the top of the

Stanwell Park Claystone was highly to completely weathered for a distance of 1 m to

4 m. The assessment that the failure plane is more likely to be located in this

weathered claystone than in the talus, is suggested by an examination of measured

physical properties of the two materials.

Since cohesion (c) in a residual strength situation is progressively reduced with small

increments of creep strain, the controlling strength parameter is generally considered to

be the residual angle of internal friction (cj)r).

The results of tests (State Rail Authority, 1982) indicate that the average ((j)r) for

undisturbed talus (27) exceeds that of the completely weathered claystone (16) by 70%.

An approximate comparison of the relative residual strength of the colluvium and the

completely weathered claystone can be made by using the relationship between plasticity

index (Ip) and residual angle of internal friction (<\>T) published by Kanji (1974):

<|)r = 46.6/Ipa446

The relevant values of Ip for talus and completely weathered claystone are 19.6% and

33.2% respectively. The equivalent values of <j)r are found to be 12.4 and 9.7. Again
217

the angle for the talus is seen to exceed that for the completely weathered claystone,

suggesting that the basal sliding surface would more likely be located within the

weathered claystone.

8.7.7.5 Conclusion

The major feature of the Coalcliff slump is sliding of the talus layer on an inclined

failure surface toward the sea, with internal tension cracks and minor scarps. As

described before the Stanwell Park Claystone is highly to completely weathered for a

distance of 1 m to 4 m. The assessment that the failure plane is more likely to be

located in this weathered claystone than in the talus was suggested by a comparison of

the measured physical properties of the two materials. Since cohesion (c) in a residual

strength situation is progressively reduced with small creep strains, the controlling

strength parameter is generally considered to be the residual angle of internal friction.

Groundwater flow appears to be introduced from an aquifer in the Stanwell Park

Claystone into the bedrock/talus interface. This could have a controlling effect on the

instability of this slope. Therefore, lowering of the groundwater within the talus and

fill may not produce the full anticipated reduction in water pressures at the failure plane

due to the relatively constant pressure applied by the flow from the aquifer below the

interface. The relative effect of the aquifer uplift pressure would appear to depend on

the surface area and the permeability of the interface materials. For example, a decrease

in the permeability of the basal talus materials (which may include completely weathered

claystone) would cause an increase in the uplift pressures on the overlying slip mass.

8.8 SURFACE SURVEY RESULTS


The writer installed in late 1991 a series of surface survey pegs on the four landslides

(Clifton hotel earth-slump, Moronga Park slump-earth flow, Jetty slump and Harbour
218

slump) for measuring slip movement. The survey pegs are arranged along three transect

lines which extend across each landslide. The survey pegs are positioned about 10 m

apart along each transect line. The landslides did not show any surface movement for

more than three years of this study. The only exception was at the Moronga Park

slump, where the rate of opening of a tension crack in Moronga Park is about 3 mm

per year on the surface. Even the Newcastle earthquake on 6 August 1994, which

measured 5.3 on the Richter scale (4.2 magnitude in the study area) had no effect on

the slumps. A monitoring program with inclinometers by the State Rail Authority

(1982) showed that the landslides are moving very slowly (creep). The rate of

movement is between 30 mm and 50 mm per year depending on the amount and rate

of rainfall in the landslide areas.

8.9 FAILURE OF ROCK


8.9.1 INTRODUCTION
The toppling mass-wasting process has been recognised throughout the world but little

is known about its long-term effect in natural rock slopes (Pritchard et al, 1990;

Culshow and Bell, 1991; Glawe et al, 1993; Cruden and Xian, 1994).

In the northern Illawarra, the Narrabeen Group is thick and slopes are steep with the

sandstone beds locally forming significant cliffs. The extreme is between Coalcliff an

Clifton where the Scarborough and Bulgo Sandstone bluffs obscure the topmost

Hawkesbury Sandstone cliff when viewed from the road. The cliff section from just

south of Clifton to Austinmer is very steep and significant toppling and falls occur i

this area. Rock failures are considered significant when a clear surface expression

results; which generally means that newly exposed faces are evident and vegetation

below the fall has been severely damaged. The rock mass involved may range in

weight from approximately one tonne to hundreds of tonnes.


219

During the present study very heavy rainfalls (700 mm per month) were not recorded

but heavy falls (400 mm per month) occurred in recent years (April 1988). This

resulted in some significant rock failure and talus failure on the escarpment as a whole,

especially between Coalcliff and Clifton along Lawrence Hargrave Drive.

A common occurrence is the flaking off of claystone fragments and after heavy rainfall

accumulations of fine claystone fragments up to 1 m deep may be seen on the sides of

road. This common feature generally goes unnoticed and only the large rock falls are

recorded. However extreme fretting of the fine-grained rock units occurs which

eventually leads to undercutting the sandstone cliffs.

8.9.2 MECHANICS OF ROCK FAILURES


The interbedded sandstone and claystone of the Narrabeen Group often produce near

vertical cliff faces which evolve by rockfall, toppling (Fig. 8.50a) and general

degradation processes. Commonly a distinct time delay occurs between heavy rainfall

and failure. This time delay varies greatly from merely a day to several months.

Rock failures on the Illawarra escarpment are relatively simple in their form and shape

because the bedding is very close to horizontal and jointing is approximately vertical.

The iron staining often extends to the base of the failure surface indicating open joint

systems at least near the surface. At the base of the massive, horizontally bedded

sandstone is a highly weathered claystone or shale. The rock failures are bounded by

joints and the basal weathered claystones or shale. Erosion of the claystone or shale

by fretting has undercut the sandstone.

In summary, several features of rock failure have been observed in the study area.

(1) Commonly, bedding is horizontal and jointing is vertical.

(2) Failures are usually associated with heavy rainfall but usually a distinct time delay

is found.
220

(3) The joint surfaces often are iron-stained, which means they are open to weathering

agents down to the base of the failures.

(4) Weathering and fretting of the basal claystone and shale causes significant

undercutting of the sandstone cliff, for example at the contact between Scarborough

Sandstone and Wombarra Shale between Clifton and Coalcliff along Lawrence

Hargrave Drive (Fig. 8.50b).

(5) Variation in mechanical properties of the weathered rocks, and the importance of

creep.

(6) High horizontal stresses parallel to the cliff faces. The vertical joints perpendicula

to the cliff face are generally closed.

8.9.3 EFFECTS OF WEATHERING AND JOINTING


The main effect which alters the properties of rocks on a slope is weathering. The rate

of weathering is variable; it is greatly dependent on the rock type and may be a short

or long term factor. The variations, usually seasonal, of temperature and rainfall can

induce alternating stresses within a rock mass which invariably reduces its strength. In

a very long-term sense, the regional stresses within an area may vary and these may

effect the stability over a period of time. Variations in the position of the water table

are often time-dependent and this causes water pressure fluctuations and various

consolidation effects in the colluvium. Variations over time in the physical and

chemical action of the groundwater also induce breakdown of the rock mass.

Rocks containing soluble minerals are particularly susceptible to breakdown over time,

as are those containing swelling clay minerals. The permeability of a rock may,

therefore, alter greatly resulting in a complex relationship of permeability and strength

between unweathered and weathered rock.

The properties of discontinuity surfaces are significantly altered with increased

weathering. The influence of joint irregularities is decreased as the strength of the joi
221

asperities is reduced (Glawe et al, 1993). This increases the possibility of them being

sheared off rather than overridden. Although the cohesion may increase due to the large

amount clay minerals produce during weathering, the friction between the joint surfaces

is likely to decrease markedly. This needs to be confirmed by scientific observation.

Patton and Deere (1970) have pointed out that weathering tends to zones of materials

with different permeabilities. A lower permeability layer overlying a higher permeability

bed can result in the formation of artesian pore water pressures in weathered slopes,

which could induce a slope failure. Weathering also is a strong undercutting agent

because of its differential action and thus it is most important in rock slope stability

studies. The toppling mechanism of failure is very often induced by undercutting

(Canuti et al, 1993).

In the study area relatively thick sandstone beds usually overlie thinner claystone beds.

The weathering of sandstone beds occurs, but the rate of weathering of the claystone

is much greater than for the sandstone. As a result, the sandstone units are relatively

fresh to slightly weathered, while the claystone units are slightly, moderately, highly or

completely weathered. The sandstone beds possess a number of vertical joints which

become less dominant back from the free face. Water movement down the joints in the

sandstone causes more intense weathering in the claystone. Water movement away from

the joints is along the contact between the sandstone and claystone causing weathering

and fretting in the claystone. Field observations of failure have shown the existence of

completely weathered to moderately weathered zones at these junctions. The intensity

and continuity of jointing, and the extent of basal weathering and undercutting determine

the size of the rock-falls and toppling. Tectonic joints are usually continuous over long

distances. The joints tend to break the rock mass into various shaped prisms. Thicker

and coarser sandstone units usually have a lower intensity of jointing than the thinner

sandstone units.
222

8.9.4 CREEP

In long term rock slope stability studies, creep is an important factor. It is well kno

that creep occurs when rock material is subjected to relatively high stresses for a lo

time. Weathered rocks especially are very susceptible to creep. Even at low stresses,

certain rock types with moderate to high moisture contents can exhibit large time

dependent strains (creep). Creep is a movement which is a consequence of erosion and

unloading at the bottom of a slope where the beds bulge upwards. Creep may occur

on wett defined planes but this is difficult to observe because it may be taken up by

an infinite number of small shear planes (Huang et al, 1994). A failure may not exist,

although progressive displacements may have been measured.

Hamel and Adams (1974) suggested that creep may be an initial or intermediate stage

in the development of toppling failures. This suggestion is important and there appears

to be a relationship between toppling and rock creep in this study. With time, there

is usually an increase in water content and a reduction in shear strength in the

claystone. The increase in water content occurs as a result of weathering with a

consequent decrease in grain to grain contact and perfection of mineral packing, and an

increased porosity. Seasonal variation in groundwater level causes a variation in shear

strength of clay. So that the time until failure is strongly dependent on the climatic

conditions.

As the clay in a slope is weathered, it is able to creep downslope if the shear stresse

in the failure zone due to gravitational forces are equal to the residual shear strengt

If the clay is highly disturbed and the strength of the failure surface has been reduc

from the peak to a residual strength, a slide may follow.

A distinction should be drawn between creep along discontinuities and within the rock

material itself. In the former case, slow gradual movements occur primarily along joint

surfaces or along bedding or failure planes. In the latter case, the intact rock materi
223

deforms at a microscopic scale under the influence of a load. Both types of creep are

likely to occur prior to a rock failure.

8.9.5 ROCKFALL AND TOPPLING ALONG THE LAWRENCE HARGRAVE


DRIVE
Along Lawrence Hargrave Drive between Clifton and Coalcliff the road is subject to

both rockfall and toppling from above and slumping of the underlying Wombarra Shale

(Fig. 8.51). Along this road many slumps are located in the talus and Wombarra Shale;

the latter generally occurs within approximately 2 m of the surface. These slumps have

been initiated by the removal of support from the Wombarra Shale due to coastal

erosion processes. The interbedded sandstone and shale sequence in the lower

Narrabeen Group are responsible for rock slope instability within the detailed study area.

Undercutting (weathering and erosion) of cliffs at the top of claystone and shale units

produces topples and rockfalls. Between Clifton and Coalcliff the Wombarra Shale is

located at the base of the Scarborough Sandstone cliff which fails due to weathering of

the basal shale (Fig. 8.51b).

The Bulgo Sandstone forms the majority of the steep escarpment in the study area. The

Stanwell Park Claystone is located at the base of Bulgo Sandstone cliff. This claystone

unit rapidly weathers and breaks d o w n to form clay typically possessing low shear

strength properties. This process is aided by the many joints in the Bulgo Sandstone

and Stanwell Park Claystone. Also it is considered that, over geological time, creep

strains would affect rock slope stability because the claystone is more likely to creep

than the sandstone unit. Therefore partings along bedding planes occur, and the Bulgo

Sandstone c o m m o n l y fails due to weathering of the basal claystone (Fig. 8.52a, b), and

in small scale due to the interbedded claystone or shale in the Bulgo Sandstone.
224

Minor rockfalls are particularly evident where the finely laminated shales are being

weathered continuously causing the more competent sandstone blocks to fall from the

cliffs between Clifton and Coalcliff.

8.9.6 ROCKFALLS AND TOPPLING ALONG THE COASTLINE


Marine undercutting takes place where the cliffs are approximately vertical and this

includes virtually all of the cliffs between Clifton and Coalcliff. Failure due to

undercutting may take one of two forms: that of toppling where the weight vector of

a block (with an approximately rectangular cross-section defined by dimensions b and

h) falls outside of the base of the block (Fig. 8.53). A master joint aUows release of

the block from the cliff and toppling occurs. Alternatively, fracturing through rock

bridges may take place, the fracture joints a line of existing joints leading to failure

(Fig. 8.53). In fact coastal erosion has undermined the Coal Cliff Sandstone leading to

rockfalls into the sea. This in turn leads to slides from the overlying Wombarra Shale

(Fig. 8.53). A notch exists at the high water mark level in the Coal Cliff Sandstone

indicating the strong coastal erosion forces. King tides combined with heavy storms

resulted in large rockfalls on the coast immediately north and south of Stanwell Park

in the Scarborough Sandstone (Fig. 8.53). In these areas the Scarborough Sandstone is

very coarse and tends to be conglomeratic. It was observed that during failure the rock

broke almost invariably at the conglomerate/sandstone interface.

8.9.7 CONCLUSIONS
There are several factors contributing to the present instability on the lower Narrabeen

Group exposed in the northern Illawarra. These include:

(1) the high rainfall levels experienced and, as a result, the high rates of infiltration

and runoff;
225

(2) the steep surface slope angle which forms a potentially unstable condition in the

Bulgo Sandstone and Scarborough Sandstone;

(3) the rapid weathering of the Stanwell Park Claystone unit which breaks d o w n to

form clay typically possessing low shear strength properties. This unit has a very

low durability (see chapter 7);

(4) the low surface slope angle on the Stanwell Park Claystone which allows the

accumulation of deep talus deposits;

(5) pre-existing discontinuities, such as shear zones and failure surfaces in the talus;

(6) blanketing of the jointed Bulgo Sandstone and any fractured basal Stanwell Park

Claystone by talus causes hydrostatic pressures to build up within the overlying

impermeable talus deposit. This is confirmed by the many failures which occur

on the Bulgo Sandstone and Stanwell Park Claystone;

(7) concentration of hydrostatic pressures within fractured rock immediately behind the

talus;

(8) the presence of interbedded claystone within the Bulgo Sandstone, Scarborough

Sandstone and Coal Cliff Sandstone which weathers to form low strength clay.

For example, claystone bands within the Bulgo Sandstone show very low durability

and cause m a n y rockfalls;

(9) The weathering of the Wombarra Shale which breaks down to form low strength

clay;

(10) the existence of seepage concentrations within the fractured Wombarra Shale;

(11) marine undercutting (Coal Cliff Sandstone and Scarborough Sandstone) especially

between Clifton and Coalcliff and the north of Stanwell Park Beach.

(12) the presence of coal seams results in seepage concentrations because of their

permeable and aquifer properties; and


226

(13) the difference in the creep properties of the two predominant rocks (sandstone and

claystone). The weathered claystone has a faster creep rate than the sandstone

(Evans, 1978).

8.10 TREATMENT, STABILISATION AND PREVENTION


Landslides are natural geomorphic processes. For natural slopes in areas where there

will be no adverse effects, landslides are simply left alone. In critical areas, landslid

on natural slopes are treated similarly to those in man-made slopes. The most critical

item here is understanding the geotechnical framework of the slope. If this framework

is understood, it is not usually difficult to implement proper design and construction

procedures, provided that the economics of such procedures can be justified.

Deep-seated rock slides are treated on an individual basis because of their rarity.

Standard procedures are available for dealing with the more common shallow, slab-

type, rockfalls and rock topples. In natural rock slopes, the options are usually limited

to removal of potentiaUy unstable rock masses or supporting and/or stabilising such rock

masses.

With talus slopes, the key is recognition of old landslide masses (Gray at al, 1979;

Hamel and Adams, 1981). Talus masses, especially the larger ones, should be avoided

to the extent practicable. If they cannot be avoided, talus masses can sometimes be

stabilised with drained buttress fills or retaining structures. Stabilisation of a talus

by excavation alone generally requires removal of virtually the entire mass in order to

ensure stability. This is seldom practical with large talus masses. Improvement of

subsurface and surface drainage is an important component of stabilisation measures for

many talus slides, though drainage by itself may not be sufficient for slide stabilisatio

(Chowdhury, 1980).

The stability of filled slopes begins with the foundation. All foundations must be

carefully investigated. Fills placed on talus are seldom stable in the long term.
227

Assuming a stable foundation, good grading practices are mandatory to ensure a stable

fill slope. Surface and subsurface drainage must also be provided for a stable fiU.

With rock slopes, individual methods of stabilisation will very rarely be used by

themselves. A realistic design will frequently involve a combination of control of

passive factors, such as weathering, the use of structural restraints, such as rock anchors,

and drainage. In m a n y cases in the study area, the prevention of rock failures in natural

rock slopes will not be viable as access is very difficult and the cost would be very

high.

A s weathering plays an important role, its minimisation as an undercutting factor in

softer rocks is most important. Physical weathering processes, especially wetting and

drying, need to be prevented as soon as possible after a fresh claystone bed is exposed

by rock failure. Detailed work on such beds is described by Fookes and Sweeney

(1976) and usually involves the sealing of the exposed bed with reinforced concrete but

with appropriate drainage requirements. Complete sealing would minimise wetting and

drying and also salt weathering processes. The rapid rate of weathering of the claystone

necessitates prompt action, especially as the strength decrease with weathering is most

pronounced in the progression from the fresh to slightly weathered stage. Large areas

of high weatherable rocks can be treated by the use of mortar screeding where

groundwater erosion is not important. Figure 8.54 shows several stabilisation and

prevention measures carried out between Coalcliff and Clifton. The Scarborough

Sandstone is at the top of the figure. The Wombarra Shale (middle) overlies the Coal

Cliff Sandstone (bottom) and the Otford Sandstone M e m b e r form the small cliff beside

the roadway. For controlling lateral movement, rockfalls, toppling and subsidence,

significant work has been carried out by the Department of Main Roads along Lawrence

Hargrave Drive including concrete retaining walls, shotcrete, gabions, rock bolts and

steel mesh to prevent falls and slumps from and onto the side of the road (Fig. 8.54).

Figure 8.55a shows destruction of the shotcrete by swelling of the rock mass. The
228

Wombarra Shale is strongly affected by wetting and drying processes which lead to the

expansion of the surface rockmass and water access is provided via the joint systems

in the overlying rock.

This study indicates the progressive nature of toppling and the general degradation

processes whereby open joints become enlarged and failure ensues. The principal

method of minimising the effects of jointed regions is the use of structural restraints i

the form of rock bolts or anchors. They should be applied as soon as possible after

exposure of the particular face so as to minimise joint dilation. The interlocking effect

of joints imparts a considerable strength to a jointed rock mass. If loosening is

permitted strength is reduced. Rock bolts or anchors act as shear keys and are designed

to provide an increased shear resistance across thinly bedded rocks.

The provision of adequate drainage is a necessary requirement of almost all stabilisation

designs. This may be in the form of subhorizontal drains drilled both paraUel and

perpendicular to the slope, or drains at the crest and toe of the slope (Fig. 8.55b). The

drains should have an impermeable lining and may act to re-channel existing permanent

flow or collect surface runoff. The amount of infiltration of surface water into the

joints at the back and sides of the slope is most critical. This can be minimised by

using drains or by sealing the cracks. The time delay noted following heavy rainfall

gives an indication of the pore pressure build-up in these joints. Subsurface

groundwater flow is an important weathering agent. This can only be avoided by

tapping the groundwater far back from the exposed face with drainage adits which will

be warranted in large scale failures.

Other methods used to control local rockfalls and general degradation are benches,

concrete or masonry retaining walls, free hanging mesh nets suspended from above,

stone facing with graded filter, rock trap ditches at the toe of the slopes, scaling of

loose blocks the construction of fences or walls and finally the flattening of the slope.

Local benches exit throughout the Narrabeen Group and greater use could be made of
229

them as they can act as access roads, rockfall arresters and can form the basis of a

contour drainage system (Fookes and Sweeney, 1976).

The coastal erosion is most critical and large breakwater walls must be constructed to

combat this process. Rock failure, both of the coastal cliffs and of the escarpment itself

must be seen as a normal phenomenon of the erosion cycle and hence most measures

to prevent them will have a definite and limited life unless they are maintained updated,

periodically and replaced.

At present, to enhance stability along Lawrence Hargrave Drive between Coalcliff and

Clifton, the following measures need to be considered:

(1) repair the rock retaining fences;

(2) seal cracks and fill depressions in the road surface;

(3) remove all loose, failed and slipping material;

(4) rockbolt, mesh and shotcrete can be applied to some locations, for example the top

of the southern and northern amphitheatre areas could be protected.

(5) replace rock fences with double strength fences;

(6) replace drainage pipes with larger sizes, and improve the inlet conditions for, and

outlet dispersal of, surface runoff water.

(7) excavate n e w drainage trenches where possible (between the cliff and the road).

(8) construction of a breakwater at base of the cliff between Clifton and Stanwell Park

along the coastline where the coastal erosion is critical.

(9) a series of deep subsoil slot drains is needed in the lower part of the landslides.

These should be taken d o w n to bedrock and extended partially under the road to

intercept the regional water table and provide adequate drainage from the

subsurface aquifer.

A n alternative to remedial work along the existing roadway would be the excavation of

a tunnel to carry traffic past the dangerous section or construct a bridge across the shore

platforms. However, such proposals m a y not be cost effective.


230
231

CHAPTER 9

SUMMARY AND CONCLUSIONS

9.1 INTRODUCTION

In the niawarra area, with its steep coastal escarpment, slope stability is an old problem

and can have disastrous effects on development. This is directly related to the geology

and geological history of the area. The 300 m high escarpment consists of flat-lying

Permo-Triassic volcaniclastic coal measures plus fluviatile sequences capped by a well

cemented quartz sandstone. The interbedded strong sandstone and weak shale succession

in the lower part of escarpment has been acted upon by erosion, stress relief,

weathering, creep and sliding processes to produce masses of talus on many of the steep

hillsides. Rockslides are rare but rockfalls and toppling are common on the steep rock

slopes in the area.

The majority of slope stability problems of economic significance in northern Illawarra

involve translational or rotational slides or slow to rapid flows of soil, talus or fill. The

precarious equilibrium of talus masses is frequently upset by heavy precipitation and by

man's activities, e.g. removal of toe support, loading the slope, and changing the surface

and subsurface drainage.

Artificially induced slides, many of which include underlying or adjacent talus, almost

invariably result from poor site selection or poor design and construction practices.

Slope stability evaluation in the northern Illawarra should be an interdisciplinary

geotechnical endeavour requiring concepts from engineering geology, soil mechanics and

rock mechanics. Of these three disciplines, engineering geology is probably the most

important. Reliable evaluations of slope stability must begin with an understanding of

regional and site geology and of the geologic processes which formed the site and

continue to act upon it. Once this level of geologic understanding is reached, slope
232

behaviour can often be assessed on the basis of common sense or precedence. Where

detailed stability analysis is required, the above-mentioned geologic understanding is

mandatory for development of appropriate geotechnical models. This study is structured

according to the above philosophy, with emphasis on the talus slope deposits.

9.2 STRATIGRAPHY
The lower Narrabeen Group and upper Illawarra Coal Measures are essentially flat-

lying strata consisting of repeated beds of sandstone, shale, claystone and coal seams.

In the lower Narrabeen Group, thick sequences of weak rocks (Stanwell Park Claystone

and Wombarra Shale) are rather more easily eroded than sandstone strata and hence

relatively rapid rates of recession occur. Undermining along this thick sequence of

weak rocks at the contact between claystone and sandstone reduces the support for the

overlying vertically-jointed sandstone and eventually leads to stabs falling off along th

vertical joint faces. Thin marker beds (coal seams) in the Illawarra Coal Measures

commonly act as aquifers, with claystone beds acting as aquitards. Slope instability is

usually related to the presence of the aquifers which are the source of high pore water

pressures.

9.3 PETROLOGY
The Narrabeen Group was derived from the New England Fold Belt to the north and

consists predominantly of volcanic detritus. The volcanic detritus is present in both the

sandstone and shale units either in form of detrital grains of volcanic rock or as fine

volcanic ash. During post-depositional alteration and diagenesis, the original volcanic

glass in the ash and matrix of larger grains has devitrified to produce smectite clays.

These clays not only cause swelling and shrinkage near the surface as a response to

wetting and drying, but also reduce the permeability of the near the surface rock mass.
233

This latter factor increases the aqueous pore pressures and hence increases the likelihood

of surficial mass movement of both the rock mass and the adjacent talus deposits.

Based on the petrological study the rock fragments mostly comprise chert grains. The

majority of the chert contains clay minerals. Coarser sandstones tend to be poorly to

very poorly sorted whereas finer sandstones tend to be moderately to well sorted. Due

to an increase in weathering from fresh to weathered rock there is: (a) a decrease in

siderite and calcite content and a resultant decrease in the percentage of the cementing

matrix with an increase in weathering; (b) an increase in overall iron staining and

increase in thickness of iron oxide coatings around the edges of the quartz grains and

rock fragments with an increase in weathering; (c) an increase in chlorite content and

diagenetic alteration to chlorite enhances the weatherability of rock fragments; and (d)

quartz and rock fragments become progressively more fractured.

Based on X-ray diffraction analyses the carbonates are mostly rare in the talus deposits.

The natural reduction in the carbonate due to weathering is one cause of talus slope

instability in the niawarra area.

9.4 STRUCTURE
The faults act as feeders for underground water and even after prolonged dry spells

water is still running from area. The net result is that a relatively small rainfall can

thoroughly saturate the talus in the fault zone, where it is already in a highly unstable

position. Joints in the Scarborough Sandstone and Bulgo Sandstone vary in spacing.

The widely spaced joints divide the Scarborough Sandstone into big rectangular blocks

while closely spaced joints divide the Bulgo Sandstone into moderate to small blocks.

As a result, more rockfalls occur in the Bulgo Sandstone than in the Scarborough

Sandstone in the study area. However, falls that occur in the Scarborough Sandstone

are usually much larger and more destructive.


234

The high horizontal stress environment known to exist in the Illawarra area is an

important factor which influences slope failure. The resulting joint strike maxima for

the lower Narrabeen Group shows that the most prominent joint set exposed at the

surface, with a direction between 005 and 025, has a significant effect on slope

stability in the study area.

9.5 WATER
Perched water tables have been found to be quite common in the study area because

the many claystone sequences within the Narrabeen Group are relatively impermeable.

Fracture permeability is the most important feature of groundwater movements with most

of the fractures occurring in areas of stress relief. It is quite obvious from studying th

rainfall figures and periods of prevalence of landslides that the most unstable periods

are those when the rainfall is above 400 mm per month.

9.6 GEOTECHNICAL PROPERTIES OF ROCK AND TALUS


A significant decrease in strength and durability was found to occur with a change in

mineralogy and an increase in weathering from fresh to weathered rocks. Moderately

and highly weathered claystone and shale in Narrabeen Group rocks have low to very

low durability; the later is dependent on their mineralogy, and especially on the type and

quantity of clay minerals present. The presence of calcite in shale will usually cause

high durability. Claystone samples interbedded in the Bulgo Sandstone show very low

durability. In contrast, claystone interbedded in the Scarborough Sandstone shows a

medium durability whereas claystone in the Coal Cliff Sandstone has a high durability.

The differences in the behaviour of samples is that slake durability is sensitive to the

abundance of clay minerals in these samples. Claystone interbedded in the Bulgo

Sandstone and the highly weathered Stanwell Park Claystone both have very low

durability. This has a significant effect on slope stability in the Bulgo Sandstone
235

especially w h e n the Stanwell Park Claystone acts as the bedrock for the talus mantle

between Clifton and Stanwell Park.

Sandstone units in the Narrabeen Group, which contain abundant clay minerals and

volcanic rock fragments, show significant strength loss on wetting. The Narrabeen

Group sandstone contains swelling clays, therefore, expansive forces in these rocks are

a mechanism contributing to strength loss. High proportions of expansive clay minerals

were detected in volcanic rock fragments (cherts) which suggest that clay softening in

the presence of water is important in controlling moisture related reduction of strength

in sandstone in the Illawarra area. In the long term, solubility and disaggregation in the

Otford Sandstone M e m b e r would cause an increase in the secondary porosity in this

sandstone. Differential subsidence of the Wombarra Shale is the result of solution and

disaggregation of the Otford Sandstone M e m b e r which is not uniform in character and,

therefore, increases slope instability in the area.

The geotechnical properties of talus most related to its stability are clay content,

plasticity index and residual friction angle. These parameters and the angle of natural

slopes show the talus is unstable in the long-term at slopes above 10-12.

9.7 SLOPE DEVELOPMENT PROCESS


Slope development processes can be simplified and generalised as follows: stream

erosion has carved longitudinal and transverse escarpment profiles reflecting local

stratigraphy, and has also removed lateral and vertical support from escarpment waUs.

Stress relief accompanying lateral support removal produced tension fractures and

bedding plane shear zones in rocks adjacent to the escarpment walls. These stress relief

features, along with stratigraphic and lithologic details, control groundwater flow in the

vicinity of the slopes; perched water tables and hillside springs are c o m m o n . Stress

relief features and related groundwater phenomena have accelerated physical and

chemical weathering of rocks on the slopes. Under these conditions, rockfalls and
236

topples are common. They occur on the natural slopes, with weathering and erosion

undercutting joint-bounded rock blocks which slump backward or topple forward

depending on the their geometry, support conditions, and applied forces which, in

addition to gravity, often include water. Rockfall volumes are typically small, ranging

from approximately 0.1 to 24 m\ Fine-grained and argillaceous rocks predominate in

the typical stratigraphic section, so the weathering products are usually silty clay or

clayey silt with rock fragments ranging in size from sand to very large boulders. As

weathering progresses, the strength of the near-surface soil and rock materials decreases

and they begin to creep or slide down the relatively steep escarpment walls under the

action of gravity and water forces. Deeper seated landsliding along bedding plane shear

zones resulting from escarpment stress relief have also occurred in certain locations.

Eventually, these processes produced mantles of talus soil and rock fragments on many

slopes. The composition, thickness and inclination of talus on a given slope reflect the

stratigraphy and erosional history of the slope. Where slopes are relatively flat, on ridge

tops or large erosional benches, residual soils have formed. Residual soils on benches

below steeper slope segments are often covered with talus.

Where erosion was intense, little or no talus has accumulated and rock strata are

exposed on steep slopes. Talus thickness can range from about 1 m to more than 10 m

but is typically in the order of 5-10 m in the study area. A maximum thickness

approaching 20 m occurs at the toes of slopes in thick sequences of weak rocks (e.g.

claystone and shale) where deep-seated landsliding has occurred or where accumulated

talus has not been removed by erosion. Such conditions are rare.

Talus soils in the northern niawarra are generally cohesive and in most cases are

fissured. The most important engineering implication of talus slope development is the

presence in talus of surfaces, or zones, along which shear strength has been reduced to,

and maintained at, residual or near residual levels by a combination of softening and

strain effects. Movement due to creep, sliding, or both, during slope development are
237

generaUy concentrated along one or more such surfaces or zones, commonly at the soil-

rock contact. Additional surfaces or zones of movement may exist at levels within the

talus, particularly where talus is thick.

A talus derived claystone is often finer grained than the original rock, having a larger

clay fraction and plasticity index. This is often found at the base of talus deposits

which commonly mantle the coastal terraces. As a result, in a landslide situation,

residual shear strengths often reflect the properties of the clays which are derived from

the basal claystone strata.

9.8 REMEDIAL WORKS


With talus slopes, the key is recognition of old landslide masses. Talus masses can be

stabiUsed with drained buttress fills or retaining structures. Stabilisation of a talus mass

by excavation alone generally requires removal of virtually the entire mass in order to

ensure stability. This is seldom practical with large talus masses. Improvement of

subsurface and surface drainage is an important component of stabilisation measures for

many talus slides, though drainage by itself may not be sufficient for slide stabilisation.

With rock slopes, a realistic design will frequently involve a combination of control of

passive factors, such as weathering, the use of structural restraints, such as rock anchors,

and drainage. They should be applied as soon as possible after exposure of the

particular face so as to minimise joint dilation. If dilation is permitted this strength

lost.

Subsurface groundwater flow in both jointed rock sequences and talus can only be

avoided by tapping the groundwater far back from the exposed face with drainage adits

which will be warranted in large scale failures.

The coastal erosion, which causes oversteepening of the cliffs, is most critical and large

breakwater walls must be constructed to combat this process.


238

An alternative to remedial work along the existing roadway would be the excavation of

a tunnel to carry traffic past the dangerous section or construct a bridge across the s

platforms. However, such proposals may not be cost effective.

9.9 CONCLUSIONS
Main causes of slope instability in the northern Illawarra region is due to:

(1) repeated beds of sandstone, shale, claystone in the Narrabeen Group;

(2) thin marker beds (coal seams) in the Illawarra Coal Measures;

(3) the presence of interbedded claystone within the Bulgo Sandstone, Scarborough

Sandstone and Coal Cliff Sandstone;

(4) the steep surface slope angle which forms a potentially unstable condition in the

Bulgo Sandstone and Scarborough Sandstone;

(5) the low surface slope angle on the Stanwell Park Claystone which aUows the

accumulation of deep talus deposits;

(6) the rapid weathering of the Stanwell Park Claystone and Wombarra Shale which

break down to form clay typically possessing low shear strength properties;

(7) the volcanic detritus in both the sandstone and shale units either in form of detri

grains of volcanic rock or as fine volcanic ash;

(8) the natural reduction in the carbonate due to weathering in the talus slopes;

(9) pre-existing discontinuities, such as shear zones and failure surfaces in the talus;

(10) the horizontal beddings, vertical joints, faults and high horizontal stress;

(11) the high rainfall levels experienced (more than 400 mm per month) and, as a

result, the high rates of infiltration and runoff;

(12) high fracture permeability and perched water tables;

(13) blanketing of the jointed Bulgo Sandstone and any fractured basal Stanwell Park

Claystone by talus causes hydrostatic pressures to build up within the overlying


239

impermeable talus deposit. This is confirmed by the m a n y failures which occur

on the Bulgo Sandstone and Stanwell Park Claystone.

(14) concentration of hydrostatic pressures within fractured rock immediately behind the

talus;

(15) the existence of seepage concentrations within the fractured Wombarra Shale and

Stanwell Park Claystone;

(16) significant decrease in strength and durability due to change in mineralogy and

increase in weathering from fresh to weathered rocks;

(17) very low durability in the claystones and shales; and

(18) marine undercutting (Coal Cliff Sandstone and Scarborough Sandstone) especially

between Clifton and Coalcliff and the north of Stanwell Park Beach.

At present, to enhance stability in the northern Illawarra region along Lawrence

Hargrave Drive between Coalcliff and Clifton, the following measures need to be

considered:

(1) repair the rock retaining fences;

(2) seal cracks and fill depressions in the road surface;

(3) remove all loose, failed and slipping material;

(4) rockbolt, mesh and shotcrete can be applied to some locations, for example the top

of the southern and northern amphitheatre areas could be protected;

(5) replace rock fences with double strength fences;

(6) replace drainage pipes with larger sizes, and improve the inlet conditions for, and

outlet dispersal of, surface runoff water;

(7) excavate n e w drainage trenches where possible (between the cliff and the road);

(8) construction of a breakwater at base of the cliff between Clifton and Stanwell Park

along the coastline where the coastal erosion is critical; and


240

(9) a series of deep subsoil slot drains is needed in the lower part of the landslides.

These should be taken down to bedrock and extended partially under the road to

intercept the regional water table and provide adequate drainage from the

subsurface aquifer.
241

REFERENCES
A d a m s o n , C.L., 1974. Geological report on the stability of the Western G u U y Site.
Report submitted to Coalcliff Collieries Pty Ltd.
Amaral, R.H., 1975. Special instability problems in the niawarra and Warringah Shire
areas of N e w South Wales. Proceeding 2nd Australian-New Zealand Conference
on Geomechanics, pp. 311-325. Institution of Engineers, Brisbane.
A S T M , 1991. Standard method for slake durability of shale and similar weak rocks.
Annual Book of ASTM Standards, 04.08, section 4, 880-882.
Attewell, P.B., and Farmer, I.W., 1976. Principles of Engineering Geology. Chapman
and HaH, London.
Bai, G.P., 1991. Petrology, diagenesis and reservoir potential of Narrabeen Group
sandstone, Sydney Basin, N e w South Wales. P h D thesis, University of Sydney,
Sydney (unpubl.).
Bamberry, W.J., 1992. Stratigraphy and sedimentation of the Late Permian Illawarra
Coal Measures, southern Sydney Basin, N e w South Wales. P h D thesis, University
of Wollongong, Wollongong (unpubl.).
Barton, M.E., 1988. The sedimentological control of bedding plane shear surfaces.
Proceedings of the 5th International Symposium on Landslides, pp. 73-76.
Balkema, Rotterdam.
Bell, D.H., and Pettinga, J.R., 1988. Bedding-control landslides in N e w Zealand soft
rock terrain. Proceedings of the 5th International Symposium on Landslides, pp.
77-83. Balkema, Rotterdam.
Bell, D.H., and Pettinga, J.R., 1984. Presentation of geological data. Proceedings of
the Symposium on Engineering for Dams and Canals, Alexandra. I P E N Z Proc.
Tech. G p - N.Z. Geomechan. Soc. 9/49: 4.1-4.33.
Bell, F.G., 1983. Engineering Properties of Soil and Rock. Butterworths, London.
Bell, F.G., Cripps, J.C., and Culshaw, M.G., 1986. A review of the engineering
behaviour of soils and rocks with respect to groundwater. In: Cripps, J.C., Bell,
F.G., and Culshaw, M.G. (eds), Groundwater in Engineering Geology, pp. 1-23.
Geological Society of London, London.
Bell, F.G., 1993. Engineering Geology. Blackwell Scientific Publications, London.
Bhandari, R.K., Senanayake, K.S., and Thayalan, N., 1991. Pitfall in the prediction
on landslide through rainfall data. Proceedings 6th International Symposium on
Landslides, pp. 887-890. Balkema, Rotterdam.
Bhandari, R.K., and Thayalan, N., 1994. Landslides and other mass movements
including failure of cuttings in residual soils of Sri Lanka. National Symposium
242

on Landslides in Sri Lanka, pp. 73-82. National Building Research Organisation,


Ministry of Housing, Colombo.
Bieniawski, Z.T., 1975. The point load test in geotechnical practice. Engineering

Geology 9, 1-11.
Blatt, H., Middleton, G., and Murray, R., 1980. Origin of Sedimentary Rocks.
Prentice-Hall, Englewood Cliffs.
Blong, R.J., 1973. A numerical classification of selected landslides of the debris slide-
avalanche-flow type. Engineering Geology 7, 99-114.
Blong, R.J., and Eyles, G.O., 1989. Landslides: extent and economic significant in
Australia, N e w Zealand and Papua N e w Guinea. In: Brabb, E.E., and Harrod, B.L.
(eds), Landslides: extent and economic significance, pp. 343-353. Balkema,
Rotterdam.
Boggs, S., 1992. Petrology of Sedimentary Rocks. Maxwell Macmillan, N e w York.
Bowles, J.E., 1992. Engineering Properties of Soils and their Measurement. McGraw-
Hill, N e w York.
B o w m a n , H.N., 1972. Natural slope stability in the City of Greater Wollongong.
Geological Survey of New South Wales Report 14, 159-222.
B o w m a n , H.N., 1974. Geology of the Wollongong, Kiama, and Robertson 1:50,000
Sheets. N e w South Wales Geological Survey, Sydney.
B o w m a n , H.N., 1980. Southern Coalfield, upper Shoalhaven Group and Illawarra Coal
Measures. New South Wales Geological Survey Bulletin 26, 116-132.
Branagan, D.F., and Packham, G.H., 1967. Field Geology of New South Wales.
Science Press, Sydney.
Branagan, D.F., 1985. A n overview of the geology of the Sydney region. In: Pells,
P.J.N, (ed.), Engineering Geology of the Sydney Region, pp. 3-46. Balkema,
Rotterdam.
Branagan, D.F., Mills, K.J., and Norman, A.R., 1988. Sydney faults: facts and
fantasies. Proceeding of the 22nd Newcastle Symposium, Advances in the Study
of the Sydney Basin, pp. 111-118. University of Newcastle, Newcastle.
Branagan, D.F., and Pedram, H., 1990. The Lapstone structure complex, N e w South
Wales. Australian Journal of Earth Sciences 37, 23-36.
Brand, E.W. 1993. Landslides in Hong Kong caused by the severs rainfall event of
8 M a y 1992. Landslide News 7, 9-11.
Broch, E., and Frankline, J.A., 1972. The point load strength test. International
Journal of Rock Mechanics and Mining Sciences 9, 669-697.
243

Brown, I.R., 1974. The stability of slopes in Tertiary sedimentary rocks of N e w


Zealand. New Zealand Geomechanics Society. Proceedings of the Symposium on
the Stability of slopes in natural ground, 7.23-7.33.
Brunsden, D., 1979. Weathering. In: Embleton, C. and Thomas, J. (eds), Process in
Geomorphology, pp. 73-129. Edward Arnold, London.
Canuti, P., Casagli, N., and Garzonio, C.A., 1993. Slope instability at a historical
site L a Verva Monastery, Italy. Landslide News 7, 11-14.
Carr, P.F., 1983. A reappraisal of the stratigraphy of the upper Shoalhaven Group and
lower niawarra Coal Measures, southern Sydney Basin, N e w South Wales.
Proceedings of the Linnean Society of New South Wales 106, 287-297.
Carson, M.A., 1969. Models of hillslope development under mass failure.
Geographical Analysis 1, 77-100.
Carson, M.A., 1972. Hillslope Form and Process. Cambridge, London.
Cascini, L., Critelli, S., Nocera, S.D., Gulla.G., and Matano, F., 1994. Weathering
and landsliding in Slia Massif gneiss (Northern Calabria, Italy). Seventh Congress
of the International Association of Engineering Geology, pp. 1613-1622. Balkema,
Rotterdam.
Chandler, R.J., 1974. Lias clay, the long-term stability of cutting slopes.
Geotechnique 24, 21-38.
Chestnut, W., 1981. Wollongong-Port Hacking 1:100,000 Sheet. Engineering
geological consideration and environmental geological hazards. Geological Survey
N e w South Wales, Report GS81/202 (unpubl).
Chowdhury, R.N., 1976. Mechanism of natural slope failures in the Greater
Wollongong area of N e w South Wales. Search 7, 396-397.
Chowdhury, R.N., 1980. Landslides as natural hazards - mechanisms and uncertainties.
Geotechnical Engineering 11, 135-180.
Chowdhury, R.N., and Young, A.R.M., 1987. Field guide to slope instability in the
Wollongong-Picton-Nattai Region of N e w South Wales. Fifth International
Conference and Field Workshop on Landslides. University of Wollongong,
Wollongong.
Cleary, J.R., and Doyle, H.A., 1962. Application of a seismograph network and
electronic computer in near earthquake studies. Bulletin Seismic Society American
52, 673-682.
Coates, D.R., 1990. The relation of subsurface water to downslope movement and
failure. In: Higgins, C.G., and Coates, D.R., (eds), Groundwater Geomorphology;
244

The Role of Subsurface Water in Earth-Surface Processes and Landforms, pp. 51-
76. The Geological Society of America.
Colback, P.S.B., and Wild, B.L., 1965. The influence of moisture content on the
compressive strength of rocks. Proceedings of the 3rd Canadian Rock Mechanics
Symposium, pp. 65-83. Toronto.
Conaghan, P.J., and Jones, J.G., 1975. The Hawkesbury Sandstone and the
Brahmaputra: a depositional model for continental sheet sandstones. Journal of
Geological Society of Australia 22, 275-283.
Conaghan, P.J., 1980. The Hawkesbury Sandstone: gross characteristics and
depositional environment. New South Wales Geological Survey Bulletin 26, 188-
253.
Connelly, M.A., 1970. A geological structural assessment at Appin Colliery with
reference to roof failure and directional mining. Proceedings of the Australasian
Institute of Mining and Metallurgy 234, 17-26.
Conolly, J.R., and Ferm, J.C., 1971. Permo-Triassic sedimentation patterns, Sydney
Basin, Australia. American Association of Petroleum Geologists Bulletin 55, 2018-
2032.
Cook, A.C., and Johnson, K.R., 1970. Early joint formation in sediments. Geological
Magazine 107, 361-368.
Crawford, E.A., Herbert, C , Taylor, G., Helby, R., Morgan, R., and Ferguson,
J., 1980. Diatremes of the Sydney Basin. New South Wales Geological Survey
Bulletin 26, 294-323.
Cripps, J.C., and Taylor, R.K., 1981. The Engineering properties of mudrocks.
Quarterly Journal of Engineering Geology 14, 325-346.
Crozier, M.J., 1986. Landslides: Causes, Consequences and Environment. Croom
Helm, London.
Cruden, D.M., and Xian-Qin Hu, 1994. Topple on underdip slopes in the Highwood
Pass, Alberta, Canada. Quarterly Journal of Engineering Geology 27, 57-68.
Culshow, M.G., and Bell, F.G., 1991. The rockfalls of James Valley, St Helena.
Proceedings of the Sixth International Symposium on Landslide, pp. 925-935.
Balkema, Rotterdam.
David, T.W.E., 1896. Summary of our present knowledge of the structure and origin
of the Blue Mountains of N e w South Wales. Journal and Proceedings of the
Royal Society of New South Wales 30, 33-69.
245

Dearman, W.R., 1974. Weathering classification in the characterisation of rock for


engineering purposes in British practice. Bulletin of the International Association
of Engineering Geology 9, 33-42.
Dearman, W.R., 1976. Weathering classification in the characterisation of rock.
Bulletin of the International Association of Engineering Geology 13, 123-127.
Dearman, W.R., and Fookes, P.G., 1974. Engineering geological mapping for civil
engineering practice in the United Kingdom. Quarterly Journal of Engineering
Geology 7, 223-256.
Deere, D.U., and Patton, F.D., 1971. Slope stability in residual soils. Proceedings
of the Fourth Panamerican Conference on Soil Mechanics and Foundation
Engineering, pp. 87-170. American Society of Civil Engineers, N e w York.
D e n h a m , D . 1976. Effects of the 1973 Picton and other earthquakes in eastern
Australia. Bureau of Mineral Resources, Geology and Geophysics, Bulletin 164.
D e n h a m , D., 1979. Earthquake hazard in Australia. Proceeding of a Symposium on
Natural Hazards in Australia, pp. 94-118. Australian Academy of Science,
Canberra.
D e n h a m , D., 1980. The 1961 Robertson earthquake - more evidence for compressive
stress in southern Australia. BMR Journal of Australian Geology and Geophysics
5, 153-156.
D e n h a m , H.A., Weeks, J., and Krayshek, C , 1981. Earthquake evidence for
compressive stress in southern Australian crust. Journal of the Geological Society
of Australia 28, 323-332.
Dick, J.C., and Shakor, A., 1992. Lithological control of mudrock durability.
Quarterly Journal of Engineering Geology 25, 31-46.
Dickinson, W . R . 1985. Interpreting provenance relation from detrital modes of
sandstones. In: Zuffa, G.G. (ed.), Provenance of Arenites, pp. 333-361. D.
Reidel, Dordrecht.
Dickinson, W.R., and Suczak, C.A., 1979. Plate tectonics and sandstone compositions.
American Association of Petroleum Geologists Bulletin 63, 2164-2182.
Dickson, T.W., and Weber, C.R., 1966. Investigation of roof fall in Oakdale State
Mine. New South Wales Geological Survey Report 44.
Dobereiner, L., and Freitas, M.H., 1986. Geotechnical properties of weak sandstones.
Geotechnique 36, 79-94.
Dolye, H.A., Clearly, J.R., and Gray, N.M., 1968. The seismicity of the Sydney
Basin. Journal of the Geological Society of Australia 15, 175-181.
246

Dumitru, T.A., Hill, K.C., Coyle, D.A., Duddy I.R., Foster, D.A., Gleadow, A.J.W.,
Green, P.F., Kohn, B.P., Laslett, G.M., and O'Sullivan, A.J., 1991. Fission
track thermochronology application to continentalriftingof south-eastern Australia.
Australian Petroleum Exploration Association Journal 31, 131-142.
Dunkerely, D.L., 1976. A study of long-term slope stability in the Sydney Basin,
Australia. Engineering Geology 10, 1-12.
Dyke, C.G., and Dobereiner, L., 1991. Evaluating the strength and deformability of
sandstones. Quarterly Journal of Engineering Geology 24, 123-134.
Early, K., and Skempton, A.W., 1972. Investigation of the landslide at Walton's
W o o d , Staffordshire. Quarterly Journal of Engineering Geology 5, 19-41.
East, T.J., 1978. Mass movement landforms in Baroom Pocket, south-east Queensland:
a study of form and process. Queensland Geographical Journal 4, 37-67.
Evans, R.S., 1978. Time dependent factors influencing the rock slope stability of the
Illawarra Escarpment, N e w South Wales. P h D thesis, University of N e w South
Wales, Sydney (unpubl.).
Evans, R.S., 1981. A n analysis of secondary toppling rock failures - the stress
redistribution method. Quarterly Journal of Engineering Geology 14, 77-86.
Everingham, I.B., Denham, D., and Greenhalgh, S.A., 1987. Surface wave
magnitudes of some early Australian earthquakes. BMR Journal of Australian
Geology and Geophysics 10, 253-259.
Fell, R., 1985. Slope stability in the Wianamatta Group. In: Pells, P.J.N.(ed.),
Engineering Geology of the Sydney Region, pp. 163-175. Balkema, Rotterdam.
Ferguson, H.F., and Hamel, J.V., 1981. Valley stress relief in flat-lying sedimentary
rocks. Proceedings of the International Symposium on Weak Rocks, pp. 1235-
1240. Balkema, Rotterdam.
Flentje, P., 1996. Assessment of slope stability in the Illawarra. P h D thesis, University
of Wollongong, Wollongong (in preparation).
Folk, R.L., 1980. Petrology of Sedimentary Rocks. Hemphill Publishing Company,
Austin.
Fookes, P.G., 1967. Site Investigation of the Siwalik Clay Bedrock, Mangla D a m .
P h D thesis, University of London, London (unpubl.).
Fookes, P.G., and Sweeney, M., 1976. Stabilisation and control of local rock falls and
degrading rock slopes. Quarterly Journal of Engineering Geology 9, 37-55.
Fookes, P.G., Gourly, C.S., and Ohikere, C , 1988. Rock weathering in engineering
time. Quarterly Journal of Engineering Geology 21, 33-57.
247

Franklin, J.A., and Chandra, R., 1972. The slake-durability test. International
Journal of Rock Mechanics and Mining Sciences 9, 325-341.
Frankline, J.A., and Dussealts, M.B., 1989. Rock Engineering. McGraw-Hill, N e w
York.
Fredrich, J., McCaffrey, R., and Denham, D., 1988. Source parameters of large
Australian earthquakes from body wave form inversion. Geophysical Journal 95,
1-13.
Gamble, J.C., 1971. Durability-plasticity classification of shales and other argillaceous
rocks. P h D thesis, University of niinois, niinois (unpubl.).
Geological Society Engineering Group Working Party, 1972. The preparation of
maps and plans in terms of engineering geology. Quarterly Journal of
Engineering Geology 5, 293-381.
Ghayoumian, J., Nakajima, S., Fatemi, S.M., and Ozaki, M., 1993. Weathering
characteristics of slate and green rocks. In: Anagnostopoulos et al (eds),
Geotechnical Engineering of Hard Soils-Soft Rocks, pp. 105-112. Balkema,
Rotterdam.
Ghafoori, M., Airey, D.W., and Carter, J.P., 1994. The durability of Ashfield Shale.
Seventh Congress of the International Association of Engineering Geology, pp.
3315-3321. Balkema, Rotterdam.
Ghobadi, M . H . and Pitsis, S.E., 1993. Notes for technical tour T T 3 - Urban slope
stability. Compiled by Chowdhury, R.N. International Conference on
Environmental Management Geo-Water and Engineering Aspects. University of
Wollongong, Wollongong.
Ghobadi, M.H., 1994. Landsliding in the northern Illawarra, N e w South Wales.
Geological Society of Australia Abstracts.
Ghobadi, M.H., 1994. Geology and slope stability in the northern Illawarra, N e w
South Wales, Australia. Seventh Congress of the International Association of
Engineering Geology, pp. 1307-1314. Balkema, Rotterdam.
Ghosh, D.K., and Srivastava, M., 1991. Point load strength: an index for classification
of rock material. Bulletin of the International Association of Engineering Geology
44, 27-33.
Gillott, J. 1989. Clay in Engineering Geology. Elsevier, Amsterdam.
Glawe, U., Zika, P., Zvelebil, M., and Rybar, J., 1993. Time prediction of a rockfall
in the Carnic Alps. Quarterly Journal of Engineering Geology 26, 185-192.
248

Gray, N.M., 1976. Geological appreciation of the seismicity of the southern portion
of the Sydney Basin. Bureau of Mineral Resources, Geology and Geophysics,
Bulletin 164, 9-10.
Gray, N.M., 1982. Direction of stress, southern Sydney Basin. Journal of the
Geological Society of Australia 29, 277-284.
Gray, R.E., Ferguson, H.F., and Hamel, J.V., 1979. Slope stability in the
Appalachian Plateau, Pennsylvania and West Virginia, U S A . In: Voight, B. (ed.),
Rockslides and Avalanches, pp. 447-471. Elsevier, Amsterdam.
Greenhalgh, S.A., Denham, D., McDougall, R., and Rynn, J.M., 1989. Intensity
relations for Australian earthquakes. Tectonophysics 166, 255-267.
Greminger, M., 1982. Experimental studies of the influence of rock anisotropy on size
and shape effects in point load testing. International Journal of Rock Mechanics
and Mining Sciences and Geomechanics, Abstract 19, 241-246.
Grice, R.H., Kim, C.S., and Brown, G.R., 1982. Relationship of texture, composition
and absorption properties to the weathering of mudrocks. Geological Survey of
Canada, Paper 82-1A, 359-367.
Grim, R.E., 1962. Applied Clay Mineralogy. McGraw-Hill, N e w York.
Grim, R.E., 1968. Clay Mineralogy. McGraw-Hill, N e w York.
Guadagno, F.M., Mele, R., and Bassi, M.T., 1994. Slope instabilities and geological
conditions in the Sant' Arcangelo Basin (Southern Italy). Seventh Congress of the
International Association of Engineering Geology, pp. 1557-1562. Balkema,
Rotterdam.
Guidicini, G., and Iwasa, O.Y., 1977. Tentative correlation between rainfall and
landslides in a humid, tropical environment. Bulletin International Association
Engineering Geology 16, 13-18.
Gunsallus, K.L., and Kulhawy, F.H., 1984. A comparative evaluation of rock strength
measures. International Journal of Rock Mechanics and Mining Sciences and
Geomechanics 21, 233-248.
Hallbauer, D.K., Nieble, C , Berard, J., R u m m e l , F., Houghton, A., Broch, E., and
Szlavin, J., 1978. Suggested methods for petrographic description, I S R M
commission on standardisation of laboratory and field tests. International Journal
of Rock Mechanics and Mining Sciences and Geomechanics, Abstracts 15, 41-
45.
Hamel, J.V., 1969. Stability of slopes in soft, altered rocks. P h D thesis, University
of Pittsburgh, Pennsylvania (unpubl.).
249

Hamel, J.V., 1980. Geology and slope stability in Western Pennsylvania. Bulletin of
the Association of Engineering Geologists 17, 1-26.
Hamel, J.V., and A d a m s , J.R., 1974. Discussion on some field examples of toppling
failure. Geotechnique 24, 691-693.
Hamel, J.V., and A d a m s , W.R., 1981. Claystone slides, interstate route 79, Pittsburgh,
Pennsylvania, U S A . Proceedings International Symposium on Weak Rocks, pp.
549-553. Balkema, Rotterdam.
Hancox, G.T., 1974. Geological aspects of slope stability. New Zealand Geomechanics
Society. Proceedings of the Symposium on the Stability of slopes in natural
ground, 4.1-4.13.
Haney, M.G., and Shakoor, A., 1994. The relationship between tensile and
compressive strengths for selected sandstones as influenced by index properties and
petrographic characteristics. Seventh Congress of the International Association of
Engineering Geology, pp. 493-500. Balkema, Rotterdam.
Hanlon, F.N., 1952. Southern Coalfield coal seams. Geological Survey New South
Wales Report 46 (unpubl.).
Hanlon, F.N., 1953. Southern Coalfield, geology of the Stanwell Park-Coledale area.
Technical Report, Department of Mines, New South Wales 1, 20-35.
Hanlon, F.N., 1958. Presidential address. Geology and transport with special
references to landslides on the near South Coast of N e w South Wales. Journal
and Proceedings of the Royal Society of New South Wales 92, 2-15.
Hardy, R., and Tucker, M., 1989. X-ray powder diffraction of sediments. In: Tucker,
M . (ed.), Technique in Sedimentology, pp. 191-228. Blackwell Scientific
Publications, London.
Harper, L.F., 1915. Geology and mineral resources of the Southern Coalfield. New
South Wales Geological Survey Memoir 7.
Hawkins, A.B., Lawrence, M.S., and Privett, K.D., 1988. Implication of weathering
on the engineering properties of the Fuller's Earth Formation. Geotechnique 38,
517-532.
Hawkins, A.B., and McConnell, B.J., 1991. Influence of geology on geomechanical
properties of sandstones. Proceedings of the International Society for Rock
Mechanics, pp. 257-260. Balkema, Rotterdam.
Hawkins, A.B., and McConnell, B.J., 1991. Sandstone as geomaterial. Quarterly
Journal of Engineering Geology 24, 135-142.
250

Hawkins, A.B., and McConnell, B.J., 1992. Sensitivity of sandstone strength and
deformability to changes in moisture content. Quarterly Journal of Engineering

Geology 25, 115-130.


Hawkins, A.B., and MacDonald, B.J., 1992. Decalcification and residual strength
reduction in Fuller's Earth Clay. Geotechnique 42, 453-464.
Hawkins, A.B., and Pinches, M., 1992. Engineering description of mudrocks.
Quarterly Journal of Engineering Geology 25, 17-30.
Helby, R., 1973. Review of Late Permian and Triassic palynology of N e w South
Wales. Geological Society of Australia Special Publication 4, 141-155.
Herbert, C , 1980a. Depositional development of the Sydney Basin. New South Wales
Geological Survey Bulletin 26, 10-52.
Hilleard, P.R., 1993. Bedrock movement and the failure of Stanwell Park Railway
via'duct. In: McNally, G.H., Knight,M., and Smith, R., (eds), Collected Case
Studies in Engineering Geology, Hydrology and Environmental Geology, pp. 1-
27. Butterfly Books, Sydney.
Hoek, H., 1971. Influence of rock structure on the stability of rock slopes.
Proceedings of the First International Conference on Stability in Open Pit Mining,
pp. 49-63. The American Institute of Mining, Metallurgical, and Petroleum
Engineers, N e w York.
Hoek, E., and Bary, J.W., 1981. Rock Slope Engineering. The Institute of Mining
and Metallurgy, London.
Hopkins, T.C, and Deen, R.C., 1984. Identification of shales. Geotech. Testing
Journal 7/1, 10-18.
Huang, R.Q., Zhang, Z.Y., and W a n g , S.T., 1994. Engineering geological studies of
creeping rock slopes. Seventh Congress of the International Association of
Engineering Geology, pp. 1401-1407. Balkema, Rotterdam.
Hudec, P.P, 1976. Development of durability tests for shales in embankment and
swamp backfills, Ontario. Ontario Ministry of Transportation and
Communications, Research Report 216.
Hutchinson, J.N., 1968. Mass movement. In Fairbridge, R.W. (ed.), Encyclopaedia
of Geomorphology (Encyclopaedia of Earth Sciences Series III), pp. 688-695.
Reinhold Pubs, N e w York.
Hutchinson, J.N., 1988. General report: morphological and geotechnical parameters of
landslides in relation to geology and hydrogeology. Proceedings of the 5th
International Symposium on Landslides, pp. 3-35. Balkema, Rotterdam.
251

Hutchinson, J.N., and Bhandari, R.K., 1971. Undrained loading: a fundamental


mechanism of mudflows and other mass movement. Geotechnique 21, 353-358.
Hutton, A.C., Ferguson, C.L., and Jones, B.G., 1990. Landslip in the northern
Illawarra Coalfield. Proceedings of the 23th Newcastle Symposium, Advances in
the Study of the Sydney Basin, pp. 37-44. University of Newcastle, Newcastle.
Hutton, A.C., Jones, B.G., Bamberry, W.J., Odins, P. and Clark, B., 1990. The
distribution and evaluation of coal in the southern Sydney Basin, N E R D D P Final
Report, Project 1019, Volume 2. University of Wollongong, Wollongong.
I S R M , 1985. Suggested method for determining point load strength. International
Journal of Rock Mechanics and Mining Sciences and Geomechanics, Abstract 22,
51-60.
Jakeman, B.L., 1980. The relationship between the formation structure and thickness
in Permo-Triassic succession of the southern coalfield, Sydney Basin, N e w South
Wales, Australia. Journal of Mathematical Geology 12, 185-212.
Jeffrey, C D . , and Shakor, A., 1990. The effects of lithological characteristics on
mudrock durability. Sixth Congress of the International Association of Engineering
Geology, pp. 3061-3066. Balkema, Rotterdam.
Jenkins, C.J., Keene, J.B., Helby, R. and Seismic cruise 112B shipboard party,
1993. Proceedings of the 25th Newcastle Symposium, Advances in the Study of
the Sydney Basin, pp. 12-21. University of Newcastle, Newcastle.
Jones, B.G., 1990. The Shoalhaven Group: implications for subsequent coal measures
deposition in the southern Sydney Basin. In: Hutton, A.C., and Depers, A.M.,
(eds), Proceedings of the Southern and Western Coalfields of Sydney Basin
Workshop 3, pp. 11-19. University of Wollongong, Wollongong.
Jones, B.G., and Rust, B.R., 1983. Massive sandstone facies in the Hawkesbury
Sandstone, a Triassic fluvial deposit near Sydney, Australia. Journal of
Sedimentary Petrology 53, 1249-1259.
Jones, J.G., and Veevers, J.J., 1983. Mesozoic origins and antecedents of Australia's
Eastern Highlands. Journal of the Geological Society of Australia 30, 305-322.
Kanji, M.A., 1974. The relationship between drained friction angles and Atterberg
limits of natural soils. Geotechnique 24, 671-674.
Keefer, D.K., 1984. Landslides caused by earthquakes. Geological Society of America
Bulletin 95, 406-421.
Keller, W., 1957. The Principles of Chemical Weathering. Lucas Bros, Columbia.
252

Kelley, D.R., 1971. Basic geological factors in landsliding and rockfalls of the
Pittsburgh region, Pennsylvania. Pennsylvanian Geological Survey, Open File
Report, 47p.
Kembla Coal & Coke Pty Ltd, 1990. Geotechnical study for underground
development, West Cliff Colliery assessment - borehole IL57. Report 253/16,
(unpubl).
Kembla Coal & Coke Pty Ltd, 1991. Geotechnical study for underground
development, West Cliff Colliery assessment - borehole IL64. Report 253/18,
(unpubl).
Kendall, P.F., and Briggs, H., 1933. The formation of rock joints and the cleat of
coal Proceedings of the Royal Society of Edinburgh 53, 164-187.
Kenney, T.C., 1967. The influence of mineral composition on the residual strength of
natural soils. Proceedings of the Geotechnical Conference, pp. 123-129. Oslo.
Kulhawy, F.H., and O'Rourke, T.D., 1981. Rock mass response and machine
performance in a full-face T B M tunnel in Rochester. Proceeding of the RETC,
San Francisco.
Lewis, D.W., and McConchie, D., 1994. Analytical Sedimentology. Chapman and
Hall, London.
Lindholm, R., 1987. A Practical Approach to Sedimentology. Unwin Hyman, London.
Longmac Associates Pty Ltd, 1991. Recent instability problems 1989-1990. Report
for State Rail Authority of N e w South Wales (unpubl).
Loughnan, F.C., 1963. A petrological study of a vertical section of the Narrabeen
Group at Helensburgh, N e w South Wales. Journal of the Geology Society of
Australia 10, 177-192.
L u m b , P., 1966. The variability of natural soils. Canadian Geotechnical Journal 3,
74-97.
Mauger, A.J., Huntington, J.F., and Cressy, J.W., 1983. Report to N E R D D C Grant
No. 81/1357. Commonwealth Scientific and Industrial Research Organisation,
Division of Mineral Physics, Sydney (unpubl).
Mayne, S.J., Nicholas, E., Bigg-Wither, A.L., Rasidi, J.S., and Raine, M.J., 1974.
Geology of the Sydney Basin - a review. Bureau of Mineral Resources, Geology
and Geophysics, Bulletin 149.
M c C u e , K., Wesson, V., and Gibson, G., 1990. The Newcastle, N e w South Wales,
earthquake of 28 December 1981. BMR Journal of Australian Geology and
Geophysics 11, 559-597.
253

McElroy, C.T., 1954. Petrology of sandstones of the southern coalfield. M S c thesis,


University of Sydney, Sydney (unpubl).
McNally,G.H., 1993. Effects of weathering on the engineering properties of the
Terrigal Formation. In: McNally, G.H., Knight, M., and Smith, R. (eds), Collected
Case Studies in Engineering Geology, Hydrology and Environmental Geology, pp.
92-101. Butterfly Book, Sydney.
Melchers, R.E. (ed.), 1990. Newcastle Earthquake Study. The Institution of Engineers,
Canberra.
Memarian, H., 1993. Fracture history of Coal Cliff Sandstone in Coalcliff area, N e w
South Wales. Proceedings of the 27th Newcastle Symposium, Advances in the
Study of the Sydney Basin, pp. 113-121. University of Newcastle, Newcastle.
Mencel, V., 1974. Engineering geological importance and possible origin of the stress
relief of the rocks of the Cordillera Blanca, Peru. Bulletin of the International
Association Engineering Geology 9, 69-74.
Merril, G.P., 1897. A Treatise on Rocks, Rock Weathering and Soil. Macmillan,
London.
Merritt, A.H., and Baecher, G.B., 1981. Site characterisation in rock engineering.
Proceedings of the 22nd Symposium Rock Mechanics, pp. 47-63. Balkema,
Rotterdam, Cambridge.
Michael-Leiba, M.O., 1989. Macroseismic effects, locations and magnitudes of some
early Tasmanian earthquakes. BMR Journal of Australian Geology and Geophysics
11, 89-99.
Midgley, H.G., 1951. Chalcedony and flint. Geological Magazine 88, 179-184.
Mitchell, J.K., 1976. Fundamentals of Soil Behaviour. John Wiley and Sons, N e w
York.
Moore, M.E., Gleadow, A.J.W., and Lovering, J.F., 1986. Thermal evolution o^
rifted continental margins: new evidence from fission tracks in basement apatite
from southern Australia. Earth and Planetary Science Letters 78, 255-270.
Morgenstern, N.R., 1967. Shear strength of stiff clay. Proceedings of the
Geotechnical Conference, pp. 59-76. Oslo.
Morris, R.C., Proctor, K.E. and Koch, M.R., 1979. Petrology and diagenesis of
deep-water sandstone. Society of Economic Palaeontologists and Mineralogists
Special Publication 26, 263-279.
Mostyn, G.R., and Adler, M.A., 1991. Design of the remedial works for the Coledale
Landslide. Proceedings of the Sixth International Symposium on Landslides, pp.
791-796. Balkema, Rotterdam.
254

Nakano, R., 1967. O n the weathering and change of Tertiary mudstone related to
landslides. Soil and Foundations 7, 11-14.
Nichols, T.C., 1980. Rebound, its nature and its effects on engineering works.
Quarterly Journal of Engineering Geology 13, 133-152.
N o r m a n , A.R., and Branagan, D.F., 1984. Sydney faults: more conundrums?
Proceedings of the 18th Newcastle Symposium, Advances in the Study of the
Sydney Basin, pp. 125-127. University of Newcastle, Newcastle.
Norris, R.M., and Bock,W., 1990. Erosion of seacliffs by groundwater. In: Higgins,
C.G., and Coates, D.R., (eds), Groundwater Geomorphology; The Role of
Subsurface Water in Earth-Surface Processes and Landforms, pp. 283-317. The
Geological Society of America.
Odins, P., Bamberry, J., Hutton, A., Jones, B., and Organ, M., 1990. Petrology of
sandstones and other rocks in the Illawarra Coal Measures. In: Hutton, A.C., and
Depers, A.M., (eds), Proceedings of the Southern and Western Coalfields of the
Sydney Basin Workshop, pp. 59-70. University of Wollongong, Wollongong.
Olivier, H.J., 1980. S o m e aspects of the influence of mineralogy and moisture
redistribution on the weathering behaviour of mudrocks. Proceedings of the
Fourth International Congress of Rock Mechanics, pp. 1-8. Balkema, Rotterdam.
Olivier, H.J., 1990. S o m e aspects of the engineering-geological properties of swelling
and slaking mudrocks. Sixth Congress of the International Association of
Engineering Geology, pp. 707-712. Balkema, Rotterdam.
Olivier, M., Bell, F.G., and Jermy, C.A., 1994. The effect of rainfall on slope failure,
with examples from the greater Durban area. Seventh Congress of the
International Association of Engineering Geology, pp. 1629-1636. Balkema,
Rotterdam.
Oilier, C D . , 1982. The great escarpment of eastern Australia: tectonic and geomorphic
significance. Journal of the Geological Society of Australia 29, 13-23.
Oilier, C D . , 1984. Weathering Geomorphology Texts. Oliver and Boyd, Edinburgh.
Patton, F.D., and Deere, D.U., 1970. Significant geologic factors in rock slope
stability. In: Van Rensburg, P.W.J, (ed.), Planning Open Pit Mines, pp. 143-
151. Johannesburg.
Patton, F.D., and Deere, D.U., 1971. Geological factors controlling slope stability in
open pit mines. Proceedings of the First International Conference on Stability in
Open Pit Mining, pp. 23-47. The American Institute of Mining, Metallurgical, and
Petroleum Engineers, N e w York.
255

Patton, F.D., and Hendron, A.J., 1974. General report on mass movements.
Proceedings of the Second International Congress of the International Association
of Engineering Geology, pp. 1-57. Sao Paulo.
Pells, P.J.N., 1985. Engineering properties of rocks in the Narrabeen Group. In: Pells,
P.J.N. (ed.), Engineering Geology of the Sydney Region, pp. 205-211. Balkema,
Rotterdam.
Pettijohn, F.J., Potter, P.E., and Siever, R., 1987. Sand and Sandstone. Springer-
Verlag, Berlin.
Pettinga, J.R., 1987a. The Waipoapoa landslide, a deep-seated complex block slide in
Tertiary weak rock flysch, southern Hawke's Bay, N e w Zealand. New Zealand
Journal of Geology and Geophysics 30, 401-414.
Pettinga, J.R., 1987b. Ponui landslide, a deep-seated wedge failure in Tertiary weak
rock flysch, southern Hawke's Bay, N e w Zealand. New Zealand Journal of
Geology and Geophysics 30, 415-430.
Pettinga, J.R., Bell, D.H., 1991. Engineering geological assessment of slope instability
for rural land-use, Hawke's Bay, N e w Zealand. Proceedings of the Sixth
International Symposium on Landslides, pp. 1467-1480. Balkema, Rotterdam.
Philbrick, S.S., 1960. Cyclic sediments and engineering geology. Proceedings of the
21st International Geological Congress, 49-63.
Picard, M.D., 1971. Classification of fine grained sedimentary rocks. Journal of
Sedimentary Petrology 41, 179-195.
Pitsis, S.E., 1992. Slope instability along the niawarra escarpment. M S c thesis,
University of N e w South Wales, Sydney (unpubl).
Potter, P.E., Maynard, J.B., and Pryor, W.A., 1979. Sedimentology of Shale.
Springer-Verlag, N e w York.
Price, N.J., 1963. The influence of geological factors on strength of coal measures
rocks. Geological Magazine 100, 428-443.
Pritchard, M.A., Savigny, K.W., and Evans, S.G., 1990. Toppling and deep-seated
landslides in natural slopes. In: Rossmanith, H.P. (ed.), Mechanics of Jointed and
Fractured Rocks, pp. 937-942. Balkema, Rotterdam.
R a m a n a , Y.V., and Gogte, P.S., 1982. Quantitative studies of weathering in
saprolitized charnockites associated with a landslip. Bulletin of the International
Association of Engineering Geology 19, 29-46.
Reiche, P., 1950. A survey of weathering processes and products. New Mexico
University Publication in Geology 3.
256

Rickwood, P.C, 1985. Igneous intrusives in the Greater Sydney Region. In: Pell,
P.J.N., (ed.), Engineering Geology of Sydney Region, pp. 215-307. Balkema,

Rotterdam.
Roberts, J., and Engel, B.A., 1987. Depositional and tectonic history of the southern
N e w England Orogen. Australian Journal of Earth Sciences 34, 1-20.
Rodrigues, J.D., and Jeremias, F.T., 1990. Assessment of rock durability through
index properties. Sixth Congress of the International Association of Engineering
Geology, pp. 3055-3060. Balkema, Rotterdam.
Russell, D.J., 1981. Controls on slake durability: the response of two Ordovician shales
in the slake durability test. Canadian Geotechnical Journal 19, 1-13.
Scheibner, E., 1987. Paleozoic development of eastern Australia in relation to the
Pacific region. In: Monger, J.W.H., and Franchetau, J. (eds), Circum-Pacific
orogenic belts and evolution of the Pacific Ocean basin. American Geophysical
Union Geodynamics Series 18, 133-165.
Scheibner, F., 1976. Explanatory notes on the Tectonic Map of New South Wales.
Scale 1: 1,000,000. Geological Survey of N e w South Wales, Sydney.
Selby, M.J., 1967. Aspects of the geomorphology of the greywacke ranges bordering
the lower and middle Waikato Basins. Earth Sciences Journal 1, 37-58.
Sharpe, C.F.S., 1938. Landslides and Related Phenomena. Colombia University Press,
N e w York.
Shaw, R.D., 1978. Seafloor spreading in the Tasman Sea: a Lord H o w e Rise-eastern
Australian reconstruction. Australian Society of Exploration Geophysicists 9, 75-
81.
Shellshear, W., 1890. O n the treatment of slips on the Illawarra railway at Stanwell
Park. Journal and Proceedings of the Royal Society of New South Wales 24, 58-
62.
Sherwin, L., and Holmes, G.G., 1986. Geology of the Wollongong and Port-Hacking
1:100,000 Sheets 9029, 9129. Geological Survey of N e w South Wales, Sydney.
Simpson, B., and Walton, G., 1970. Clay mylonites in the English Coal Measures:
their significance in opencast slope stability. Proceedings of the 1st International
Congress of Engineering Geology, pp. 1388-1393.
Skempton, A.W., 1953. Soil mechanics in relation to geology. Proceedings of the
Yorkshire Geological Society 29, 33-62.
Skempton, A.W., 1964. Long-term stability of clay slopes. Geotechnique 14, 77-
101.
257

Skempton, A.W., and Hutchinson, J.N., 1969. Stability of natural slopes and
embankment foundations. State-of-the Art Report. 7th Int. Conf. Soil Mech. &
Foundation Eng. Mexico, State-of-the-Art Vol, 290-340
Standard, J.C, 1969. The Sydney Basin, Triassic system, Hawkesbury Sandstone.
Journal of the Geological Society of Australia 16, 407-417.
State Rail Authority of N e w South Wales. 1982. Geotechnical report and
recommendations on stabilisation of the Coalcliff slip. State Rail Authority,
Sydney.
State Rail Authority of N e w South Wales, 1983. Geotechnical progress report and
remedial works in the Coalcliff tunnel. State Rail Authority, Sydney.
Stone, I.J., 1990. Geological assessment of coal mine roof conditions, southern Sydney
Basin. P h D thesis, University of Wollongong, Wollongong (unpubl).
Surendra, M., and Lovell, C.W., 1984. Estimation of clay minerals in clay shales by
X-ray diffraction techniques. Fifth International Conference on Expansive Soils,
pp. 27-31. The Institution of Engineers, Australia, National Conference Publication
No.84/3, Adelaide.
Taylor, R.K., and Spears, D.H., 1970. The breakdown of coal measures rocks.
International Journal of Rock Mechanics and Mining Sciences 7, 481-501.
Taylor, R.K., and Cripps, J.C, 1984. Mineralogical controls of volume change. In:
Arte w e 11, P.B., and Taylor, R.K. (eds), Ground Movements and their Effects on
Structures. Surrey University Press.
Terzaghi, K., 1936. The stability of slopes in natural clay. Proceedings of the 1st
International Conference on Soil Mechanics and Foundation Engineering, pp. 161-
165. Cambridge.
Terzaghi, K., 1950. Mechanism of landslides. Geological Society of America, Berkley
Volume, 83-123.
Terzaghi, K., and Peck, R.B., 1967. Soil Mechanics in Engineering Practice. John
Wiley and Sons, N e w York.
Trotter, C M . , 1993. Weathering and regolith properties at an earthflow site.
Quarterly Journal of Engineering Geology 26, 163-178.
Tsidzi, K.E.N., 1991. Point load - uniaxial compressive strength correlation. Seventh
International Congress on Rock Mechanics, pp. 637-639. Balkema, Rotterdam.
Turk, N., and Dearman, W.R., 1986. Influence of water on engineering properties of
weathered rocks. In: Cripps, J.C, Bell, F.G., and Culshaw, M . G . (eds),
Groundwater in Engineering Geology, pp. 131-138. The Geological Society of
London, London.
258

V a n Eechhout, E.M., 1976. The mechanisms of strength reduction due to moisture in


coal mine shales. International Journal of Rock Mechanics and Mining Sciences
13, 61-67.
Varnes, D.J., 1958. Landslide type and processes. In: Landslides and Engineering
practice. Highway Research Board, Washington, Special Report 29, 20-47.
Varnes, D.J., 1978. Slope movement types and processes. In: Schuster, R.L., and
Krizek, R.J. (eds), Landslides, Analysis and Control. National Academy of
Sciences, Washington, Special Report 176, 11-35.
Vaunat, J., 1994. Slope movement: A geotechnical perspective. Seventh Congress of
the International Association of Engineering Geology, pp. 1637-1646. Balkema,
Rotterdam.
Veevers, J.J., Powell, C M c A . , and Roots, S.R., 1991. Review of seafloor spreading
around Australia. I. Synthesis of the patterns of spreading. Journal of the
Geological Society of Australia 38, 373-389.
Veevers, J.J., and Li, Z.X., 1991. Review of sea floor spreading around Australia. H.
Marine magnetic anomaly modelling. Australia Journal of Earth Sciences 38, 391-
408.
Voight, B., 1973. Correlation between Atterberg plasticity limits and residual shear
strength of natural soils. Geotechnique 23, 265-267.
Walker, B.F., Amaral, B., and MacGregor, J.P., 1987. Slope instability in the
Coledale area of the Illawarra Escarpment. In: Walker, B.F, and Fell, R. (eds),
Soil Slope Instability and Stabilisation, pp. 417-430. Balkema, Rotterdam.
Walker, P.H., 1960. A reconnaissance of the soils and land use of part of the South
Coast of N e w South Wales. Soils and Land Use Series 38. C S I R O Division of
Soils, Sydney.
Wallis, R.G., and Johnson, M., 1969. Hydrogeological study of the Sydney Basin.
Geological Survey New South Wales, Progress Report 11, 23-34.
Walton, R.J., Enever, J.R., and Windsor, C.R., 1990. Stress regime in the Sydney
Basin and its implications for excavation design and construction. The Institution
of Engineers, Australia, Tunnelling Conference, Sydney.
Waltz, J.P., 1971. A n analysis of selected landslides in Almeda and Contracosta
Counties, California. Bulletin of the Association Engineering Geologists 8, 153-
163.
Ward, C.R., 1971a. Mesozoic sedimentation and structure in the southern part of the
Sydney Basin: Narrabeen Group. P h D thesis, University of N e w South Wales,
Sydney (unpubl).
259

W a r d , C.R., 1971b. Mineralogical change as marker horizons for stratigraphic


correlation in the Narrabeen Group of the Sydney Basin, N e w South Wales.
Journal and Proceedings of the Royal Society of New South Wales 104, 77-88.
W a r d , C.R., 1972. Sedimentation in the Narrabeen Group, southern Sydney Basin,
N e w South Wales. Journal of the Geological Society of Australia 19, 393-409.
W a r d , C.R., 1980. Notes on the Bulgo Sandstone and Bald Hill Claystone. New
South Wales Geological Survey Bulletin 26, 178-186.
Weissel, J.K., and Hayes, D.E., 1977. Evolution of the Tasman Sea reappraised.
Earth and Planetary Science Letters 36, 77-84.
Wellman, H.W., and Wilson, A.T., 1965. Salt weathering, a neglected geological
erosive agent in coastal and arid environments. Nature 205, 1097-1098.
Wellman, P., and McDougall, L., 1974. Potassium - argon ages on Cainozoic volcanic
rocks of N e w South Wales. Journal of the Geological Society of Australia 21,
247-272.
West, L.J., 1994. The Scarborough landslide. Quarterly Journal of Engineering
Geology 27, 3-6.
Wilson, R.G., Wright, E.A., Taylor, B.L., and Probert, D.H., 1958. Review of the
geology of the Southern Coalfield, N e w South Wales. Proceedings of the
Australasian Institute of Mining and Metallurgy 187, 81-104.
Wilson, S.D., 1959. Application of the principles of soil mechanics to open pit mining.
Quarterly Journal of the Colorado School of Mines 54.
Winters, D.M., 1972. Pittsburgh red beds: stratigraphy and slope stability in Allegheny
County, Pennsylvania. MSc thesis, University of Pittsburgh, Pennsylvania
(unpubl).
Young, A.R.M., 1976. The distribution, characteristics and stability of debris mantled
slopes in the northern Wollongong. M S c thesis, University of Wollongong,
Wollongong (unpubl).
Young, A.R.M., 1977. Characteristics and origin of coarse debris deposits near
Wollongong, N S W , Australia. Catena 4, 289-307.
Young, A.R.M., 1978. Influence of debris mantle and local climatic variations on
slope stability near Wollongong, Australia. Catena 5, 95-107.
Zaruba, Q., and Mencel, V., 1969. Landslides and Their Control. Elsevier,
Amsterdam.
FIGURES TO CHAPTER 1
Newcastle
1887,1889
1950.1956
1985.1972-75

Tamar Valley
1956 70

Fig 1.1 Major areas affected by landslides in Australia. N u m b e r refer to important years of
rainfall-induced landsliding (after Blong and Eyles, 1989).

Wollongong

0 150 300

Kilometres

Fig 1.2 Location of the Illawarra area.


To Sydney

HELENSBURGH

LILYVALE

OTFORDJ 1^

STANWELL PARK

/SCARBOROUGH

'WOMBARRA

'<*
<f
SOUTH
THIRROUL,

BULLI PACIFIC OCEAN

'WOONONA
Scale 1:125.000
(
Si
c I

Freeway F6
CORRIMAL Trunk Road
/ Other Main Road
Railway, Station -t 1-*-
/

Main Shopping Centre

ffl^ Harbour, Major, Other AA


Hospital. Major, Other Main
*
University, Technical College
WOLLONGONG

Fig 1.3 Location of northern Illawarra area.


150550d 1510000
3414W 3414'00

341800 |3418'00'
150550p 15100'00'
Fie 1.4 Topography of the specific area chosen for study showing the steep slope on the lower
part of the escarpment and the lack of coastal plan.
(W-E)

Hawkesbury Sandstone

Newport Formation
Bald Hill Claystone

Bulgo Sandstone

Stanwell Park Claystone


Scarborough Sandstone
Wombarra Shale
Coal Cliff Sandstone
Illawarra Coal Measures Bulli Coal

Fig 1.5 Schematic cross-section through escarpment showing geology and location of instability
between Scarborough and Clifton
/

a. Rotational failure in
overburden soil, waste
b. Plane failure in rock
rock or heavily
with highly ordered
fractured rock with no
structure such as slate
identifiable structural
pattern.

W e d g e failure on two d. Toppling failure in


intersecting discontinuities. hard rock which can
form columnar structure
separated by steeply
dipping discontinuities.

Fig 1.6 M a i n types of slope failure (after H o e k and Bray, 1981).


Crown
sceniic Cracxs

Transverse Cracxs

Degree ol
rotation

\ Transverse ~J
W ^ , C'acxs ^J
'one

rOOl

3aaiai Cracks
Toe

Fig 1.7 Definition of landslide terminology


(modified after Hansen, 1984).

FLOW
SLIDE

SLUMP CR
ROTATIONAL SLIDE
^ivxv'.v^y
ZM^^^S
FALL

M
Fig 1.8 A simple classification of landslides
(after Blong, 1973).
TYPE O F MATERIAL
TYPE O F M O V E M E N T ENGINEERING SOILS
BEDROCK
Predominantly fine Predominantly coarse

FALLS Eann tail Deoris tall R O C K fall

TOPPLES Earth topple DePris topple ROCK topple

ROTATIONAL FEW Earth slump Debris slump RocK slump


UNITS
SLIDES Earth bloc* slide Deoris block slide Rock blocK slide
TRANSLATIONAL
MANY
Earth slide Deoris slide R O C K slide
UNITS
LATERAL S P R E A D S Earm spread Deoris spread R O C K spread

Eartn How Deoris (low Rock How


FLOWS
(soil creep) Ideeo creep 1
COMPLEX Comomation ol two or more principal types ot movemeni

Fig 1.9 Varnes' (1978) classification system.

Hare substratum

e) MULTIPLE

Fig 1.10 Selected landslide types in Varnes'(1978) classification


system.
Fig 1.11 Possible large differences in fluid pressures in adjacent rock joints (after Patton and
Deere, 1971).

f f S S

a) Porous Soil Slop* b) Low Porosity Rock Slop*

Fig 1.12 Comparison of groundwater fluctuation between soil and rock slopes (after Patton and
Deere, 1971).
FIGURES TO CHAPTER 2
N
Tamworth

Bathurst

Central W e s t
Fold Belt
The Illawarra Region

Quaternary sediments
Ulladulla

Tertiary sediments
v
ii-

Triassic sediments

0 60 100 km %':::\\ Permian sediments


i_ _L_ *""/.N and volcanics

Fig 2.1 Generalised distribution of sequences in the Sydney Basin (after Young, 1976).
N

REFERENCE

~ Unconformity

Area containing volcanic:


' within the section
*<
"L^WORONORA
Monocline showing
V ^PLATEAU
rW~ycoox)B*. o direction and dip

Fault showing
yj>N, ->v<V.t Wollongong direction and dip

c^ Thrust showing
& direction and dip

40 k m Boundary of basin

.SASSAFRAS'. f j Position uncertain

Boundary of a
structural subdivision

Syncline

Anticline

Fig 2.2 Structural subdivisions in the Sydney Basin (after Branagan, 1985).
TN

Fig 2.3 Major joint trends in the Sydney Basin (after Branagan, 1985).
N

Wollongong
1200

^\. Port Kembla

Fig 2.4 Average annual rainfall (mm) in the study area (source : Bureau of Meteorology
records).
W NE
BULLI COAL
Loddon Sandstone Member
Burragorang Balgownie Coal Member
Tuffaceous Claystone Member- Lawrence Sandstone Member ECKERSLEY
Cape Horn Coal Member FORMATION
Hargrave Coal Member
Woronora Coal Member
Novice Sandstone Member
Farmborough Claystone Member-' SYDNEY
WONGAWILLI COAL
KEMBLA SANOSTONE SUBGROUP
American Creek Coal Member ALLANS CREEK
FORMATION
DARKES FOREST SANDSTONE

BARGO
CLAYSTONE Huntley Tuffaceous Claystone Member
Austinmer Sandstone Member
TONGARRA COAL
BLACKMANS FLAT CONGLOMERATE WILTON FORMATCON
Womtmn Ton I Mpmber
ERINS VALE FORMATION
Figtree Coal Member
CUMBERLAND
MARRANGAROO CONGLOMERATE PHEASANTS NEST

w
***
* Unanderra Coal Member FORMATION
Berkeley. Five Islands, Calderwood & Minumurra
Latite Members
SUBGROUP

Fig 2.5 Idealised stratigraphy of the Illawarra Coal Measures (after Odins et al, 1990).
Section Lower Half - Wombarra Shale
Coalcliff Adit
Measured Section W 308
W 606135

Mudstone,mid grey

' ' T\ - Sandstone, mudstone interbedded


J m
in units to several cm,60:40
31m Mudstone,mid grey
23 Sandstone,fine-grained, limey,
variable thickness

.91 m Mudstonetmid grey,poorly bedded

Sandstone, fine-grained,limey,
.08 m
light grey, variable thickness
.53 m Mudstone,grey
Sandstone, fine-grained,variable
.09 m
thickness, limey
.31 m Mudstone\grey

Q, Interbedded thin limey lithic


m
sandstones & mudstone,50:50.

.46 m Sandstone^edium-grainedjlithic

1.22 m Mudstone,light grey, planar base

Sandstone, medium-grained, quartz


lithic, light grey

Coal Cliff Sandstone

5923

Fig 2.6 Section of the lower half of the Wombarra Shale in the Coalcliff area (after Bowman,
1974).
Measured section from the upper Bald Hill Claystone to the Hawkesbury Sandstone
Location: Near Otford Station on Lawrence Hargrove Drive
No. W 172
Wollongong 1 : 250,000. (034226!

Sandstone, fine-grained, light yellow grey,


quartz lithic, thinly flat bedded. Typical
of normal Hawkesbury Sandstone. S o m e red
ironstained bands, interbedded with thin less
than ,03m hands of silty sandstone.
Sample VV172H.

I ,qm Silty shale, light grey


Sample W 1 7 2 G (.58m from base).

'lAn Sandstone, fine-grained off white, quartz


lithic. very hard, appears to have a chemical
cement(?quartz), grades to light grey towards
base. Appears to thicken and thin.
Sample W 1 7 2 F .

|5?m Siltstone, light grey, micaceous.


S o m e plant fossils, gradation base.
Sample W 1 7 2 E (.30m from above).

l.ooim Shale, m u d grey,fissile.S o m e micaceous


flakes.
Sample W 1 7 2 D (.40m from base).

Claystone, light grey to mid grey with


a reddish tinge, limonile lined pits c o m m o n ,
strongly outcropping claystone lends to be
flakey. Quite distinct from massive Garie
Tonstein although it tends to grade into it.
Sample W 1 7 2 C (.30m from base).

1.7J5m Tonstein. Light olive grey occasional


ironstone bands (weathering?).
Sample W 1 7 2 B .

Unconformity

Chocolate shale occasionally clay pellet.


S o m e limonite pits away from clay pellet
zones. S o m e mottling throughout section
in zones which appear continuous over lm
or more. Mottled contact with overlying
unit. Not base of claystone.
Sample W 1 7 2 A . 5628

Fig 2.7 Measured section from the upper Bald Hill Claystone to the Hawkesbury Sandstone
above the Lawrence Hargrave Drive near Otford station (after B o w m a n , 1974).
CD
O
CD

n>
L.
.n
CD m
C QJ CD CD
C c '
L-
:s
u o o c
0o
3
c ~
c
o c >. c
:> < c/> _J

j?X <tuj > 5


/ h-<o3 So
rro
' 2
500 rrO
OC <J
= 2
SO

y
* -.
- - '. -

* -

-*'-- r' \ -
* *.
- \
-
,-A, _.
I -
*
C/^^^B "

^fcVfc.
^^^^^^^^^^^^^^^^H

Fig 2.9 A series of small en echelon faults in the Jetty Fault suggests that this fault have been
active during deposition.

Fig 2 10 Rose diagram showing the distribution of forty faults in the study area (after B o w m a n ,
1974).
::::A\V".::::V/.*'.:.:!:::';': Scarborough Sandstone

Otford Sandstone
Member
> W o m b a r r a Shale

'{:'{?. Coal Cliff Sandstone


Bulli Coal

15 30 metres
metres _j i

90-
Vertical distance
60
above Bulli "
Coal
30 J

3 6 9 metres
Vertical displacement of faults

Fig 2.11 T h e Jetty Fault between Clifton and Coalcliff (after Hanlon, 1953).

Fig 2 12 Generalised diagram showing the relationship of joints and fold axes in the study area
(after B o w m a n , 1974).
\

FIGURES TO CHAPTER 3
Fig. 3.1 The basal formation of the Sydney Sub-group is the Wilton Formation which, in this
area, consists of an upward coarsening interdistributary bay fill facies overlain by the
delta top Tongarra Coal. The continuous shaly bands in the coal represent beds of
volcanic ash. The overlying Bargo Claystone strata with the 1.5 m thick Austinmer
Sandstone M e m b e r comprises coarse sandstone fining up to siltstone. South of
Wombarra Beach.

Fig. 3.2 The Bulli Coal can be seen along the waters edge from the Jetty Fault to Clifton
Fault. At the top of the photo Coal Cliff Sandstone overlies Bulli Coal and is in turn
overlain by Wombarra Shale at the top right of the photo. The laminated overbank
unit in the Coal Cliff Sandstone is truncated by erosion at the base of the second
channel-fill sequence.

Fig. 3.3 A section of the Narrabeen Group at Coalcliff. At the base of the section the Coal
Cliff Sandstone (CSs) erosionally overlies the Bulli Coal. The W o m b a r r a Shale
(WSh), on which Lawrence Hargrave Drive is built, overlies the Coal Cliff Sandstone
and the Otford Sandstone M e m b e r ( O S M ) is above the roadway. The Scarborough
Sandstone (SSs) forms the next cliff which is overlain toward the top of the photo
by Stanwell Park Claystone (SPC). The Bulgo Sandstone (BSs) forms the cliff at the
top of the photo. It is down-thrown on the right of the photo by Fault 1. At the
top left of the photo the Hawkesbury Sandstone forms top of the escarpment.
Exposed joint surfaces on the Coal Cliff Sandstone and Scarborough Sandstone show
the locations of the recent rockfalls.

Fig. 3.4 Wombarra Shale is composed of greenish grey shale with fine-grained lithic sandstone
interbeds. In the detailed study area, the unit is subject to weathering and erosion,
with active marine erosion occurring along the coastal cliff-line. The weathering of
the Wombarra Shale, which breaks down to form low strength clay, and the existence
of seepage concentrations within the fractured Wombarra Shale cause instability and
differential erosion along the coastal cliffs just south of Coalcliff adit.

Fig. 3.5 The Otford Sandstone M e m b e r has a m a x i m u m thickness of 7 metres with cross-
bedding, erosional scours, conglomeratic lenses and a few siltstone interbeds.
Undercutting is evident at the contact between the conglomeratic portion and the
cross-bedded fine-grained Otford Sandstone Member. Also the rapid weathering of
the interbedded siltstone, which breaks down to form clay, typically possesses low
shear strength properties and causes, small rockfalls. Lawrence Hargrave Drive
between Clifton and Coalcliff.
Fig. 3.3

Fig. 3.5
umi

(D
in X! CD CD
O X
^ / aO \3
c^
J3
CU G S
00 00
a. 8 c
c X!
<+-<
CD 13
z^^\

c CD <*-> X~"'
o G x
G
ea <>O/> >>
Bj
^ ocT
^
o
o o 43 *->
'"'/ / L ^ T to *& ' o "o C/3
J fH 0) x
T! "O t3 X!
* ' \ \ i_< MM Mi s I. 5T. -* T3 BO 00
^\1 CD G 3
M*m a
1c Z
4->
8 s G
O
2*. l l
C/3

a fH O 0 f0 0
' ,] CD o c/3
S CD
J(4
W l j__ X!
C/5
o
a o
w
3 aG H-i
<4-<
^^^^^B cD
60 1H
o
.s 8 *-> 00 0
13
OH 00
s
CD
4) CD
T3 X!
H X U
8 G
0
l-i

5
j3 3 >>
13 &0 X<D
^c
G T3 o cat CD
B8
CD CD X B
C 60
CD
CD
o
&0
CD CD >
c
-a
4-*
c
00

3 Qa?
CD
oo 0
C/3
X
s
s*
o
"O
2
>
U
00 G c/3 c 00
3 o
s u Xi 0 la
1-1

o
o CD 6073 CD
13 3 0 G
o s-
Cj
1
C/3
<D 00
p
O CD
S3 >
XI
o Xo O (4-1

H GO
vq
fri
00
i-l

PLH
N

{
\< unstable
area

Rnbl Bulgo Sandstone


50 m RnspiStanwell Park Claystone
I Rns (Scarborough Sandstone
Rnw j Wombarra Shale
Rnc |coal Cliff Sandstone
Bulli Coal
1 pjy \illawarra Coal Measures

Fig. 3.7 South from the Coalcliff tunnel to the Clifton Fault, the Scarborough Sandstone and
Wombarra Shale form the foundation for the railway line in a very unstable area
(Rube Hargrave Park).
CD

If
1-1
CD
>
<4H
O
GO
*H G
40
O
G
O <4H
iI

c a
S o.
xl

G ed
O 00
GO G

"S s
s
s*
s
8J?
3 S
oo G
t-l - T H
cd X

sf
<D
H CD
""I O
CD ^1
00
N
9 NORTH CLIFF i
BOREHOLE IL60


BOREHOLE 1L57
WEST CLIFF

0
BOREHOLE IL64
e
BOREHOLE IL55

( COALCUFF

1000 m

(a)

(W-E)

[L55
IL57
IL60 Out crop
(Coalcliff)

|2f)

Conglomerate

Claystone

Shale

Sandstone

T Illawarra Coal Measure

Fig. 3.11 (a) The location of the boreholes IL55, IL57, IL60 and IL64 in the West Cliff area;
(b) lower Narrabeen Group cross-section in the West Cliff area.

Hole name Easting Northing


10064 285620.30 1211500.00
10060 281980.10 1211956.30
10057 286299.70 1213674.70
10055 289501.20 1210778.60
IL60
IL57
491 453 1L64
Bulli Coal
Bulli Coal
u
Bulli Coal
I
t
;;:' '/ :'.':''; . ' . - . I

Loddon Sandstone
c
E
o
-<
C/)
T3
C
c

Bilk
on
c
C3

Balgowni sSil

p
Balgownie Coal Balgownie Coal
o yf f ra e
f m c
a ~~A
497
cu Siltstooc

-Jc
c (c)
w

(a) a
c
CO
2 m ca
u
c
S
Cape Horn seam
ed

"cf
Hargrave seam

CJ
W
CO
o

537

(b)

Fig. 3.12 Upper Illawarra Coal Measures (ICM) cross-section in the West Cliff area, (a)
borehole IL60, (b) borehole IL57, and (c) borehole IL64.
tl

m
> .
I
II iii
id

CNl *

en -i r n

*
*

_eu
. C

O
- X

1 1 P
J : ;

u
$

I
=
L
Hi in

:-y::/s .V
III II
lllll'l& :iii' ,i -;.,'.:
I'IVI'
U
CO u
rN
*
-UI n. in
*
5 "3
= "3

c TT
vC
a) _J
c/5
U)
"1)
*
UI
C O
u u
a) o
r" JC
c
U5
-n u
a -
i/: r
m
C_i ,1
0)
u c
C3 !U
o n
r
u
m
OO ui 9 1 ^H

rn
eb
u_

Ov a* Sol

?*
00
>
a*
0*
a
o
0
r

0*
o* i
oo 00
ft ' ' - "
% a
0*
T
' "I

J
fta o
*0* 0
.0
V * 0 o 1'
a
. ' - . i*i/.v.,r'i'i,i,iVi.v.,,'iv/1tl',,::/;:.\t,i,i|i,ii,1,1,1'.,,'/1'1i<'.-i:;<:.:/.':.- .-

-m c-9:
Fig. 3.14 W o m b a r r a Shale consists of mostly shale and claystone with interbedded sandstone
and conglomerate occurring locally (Otford Sandstone Member). Shale is thinly
bedded with occasional scow surface and very fine-grained sandy lenses and race
interbedded breccia - probably proximal floodplain. Marked erosional sew at the base
of conglomerate sandstone with cross-bedding evident higher in the core.

Fie 3 15 Fissilty in the Wombarra Shale has effected the mechanical properties of the unit.
T h e presence of the fissilty, unstable cements and swelling clays contribute to rapid
weathering and disintegration and reduce the overall rock mass quality in the
W o m b a r r a Shale in outcrop.
v

>n co
| | l l f
J(!i!i
in
l '!i'!ii l ' l i iiifriiJ!iiii'i i . iii!l|li:i!iiV!ii,.l!',,i.'.|1|!i.,i,,,1,i '.'.i i' i
I'liti'm
c
XI
u
u
m m
c
J3
CN C-Q en ,
,
XI s
o
ovi;
c J
H-"
CN 93
c
J=
fl)
u
O
.C
-N

sed
^

cd
9J

'_ m
CO
u -J
J
4J
CO C

2 IVl'iMi'ri'il- co "oo
JE X
c o
c s->
c c
o cd
a) rr
u:
U3 ^c
O _J
C
CN r^l t
m .. CO
co
si . I C- c/3 J2
o
l.'.'M cd
3 IIVI'.'I'I'IVI'I'I'I'I'.'I'I1. i.'ii ,1,1,1;,;[;[; I I| I 1 I||,I,I,I,I,I,III,|,I,V,I|I|I,'IV,' I ',I|'I'|| I 'I I I'I'I I I'I I I'I'I I I'I'|I I 'I I if'ii.it'r'.T-frrr
fc o
Xcd
Ec r
IF .HI c"6f-

in

VO
*
cn
D
fc

Fl
O
>0
= 2 I'I'I'IVI'IV I'm I'.'iV.'i'rVriV,' 'Yl'.',1 Wi 1:1,' l',Vl' I'lVl'

cn
m 92 in
cn -3-
"3" I

i
0=>o o.o o
r>o oo0 -a
U
0 ?0 OO s
u
^ OOO ooo
0 0 , A'' Q it
r>o . YV G
o o o
0 0 " v ' .
oO
o
-oo 1 T3
000 ' V
no .'-. -L.v nO
0 ''/[:.
00 " . ' Oo(
: % ' ,

OQO '' \
000

V
.oO :'.'.''.' 1 ':' 8 o
I Oa . T) O
II
m
CN
ui oe "o
o
I
O
in
cd

a
u

U
CD
^ o
u -
cn CN -o
o UJ 6- CN e w
3- -a
c c
o rt
a
CD VO
Oop So S
c/3 II
U
S <u
060 oo
_ Qo0 Si g
*
\
0 0 o,? B
n
Cfl
5
* , 6
CD
2C X>
."-
". . ; ; .
> o:
*. -1 cd
. ' " ' O o . O o
. " ' '
,
booo
0 . u0
in . ' . " . "
- O o oo1I
"O
-J
'.' .; . . ' - . " . ' . ' ' ;.*' . O O
' 00 3 CD
-' : oo O

in
" ' . . I '.
' . - ' " .
*
- .- . . . . . .
1 0 oO OOQ
' oo
O O
Cd _!
o o
in -UI C-Tfr. t>0 JQ
cn r-
ii

rn
00
tLi
c
0
~u*1
> 0 ,
0 5
(1i 0
n
> a
o.
a
^
3
0 n
.* ) 0 .i

/
^''- ^n . *
<

> ' .
. i

>':.
;.. i
. '

"'.
/\ ,

'
.; '
1

o
0
; ,'. . n
r
oo

'
.
I *
>
tf
Q1
0
"Q.?
B
5
3
It3
. * _
* > .* 0
o >
; ,<>...
\ * * .
o
0


I .'.

S-
0
-J . 3 ". "
a0 .'**. ft
s '- ':' Cl
* .' > .'. * , .
r, o"?.' ' .

* * I o '.'..'
I--. . . ou rB
oft
. *
^>. *

ft .'.
cn
cn
0 -* -"*
Fig. 3.18 T h e Scarborough Sandstone in borehole IL55. The photo presents the middle part
of formation consisting of fine- to very coarse-grained sandstone interbedded with fine
siltstone and conglomerate. A prominent single vertical joint with a length about 1 m is
present in the sandstone.
^1 A
- -3 O
so
} T3 J
'.' .'; V' JD
'-". "o
CD
>n 7" ,, O
in I'llll'i'l1 I1!1 i'X'X'lvll. I1 !':{'.. . .v. !:!! I'!'!'!';'!-!
-J
cn n cd
cn in g - 6i- CN CO
cn in cd
cn
t! n
""1 in
CN _J
U
oc co
<D n
g

-1
CD 13
n
J3
,.N
a
- '

C a
U O
CD cd
SO CD r,
:''.'::'"'.. :I!I!.!.!I!'" oo Tj"
SO
00 -J
1H
O
CN CD CD
SO
m m z\ cn
CD O
cn O D
00
c
>...
cd
O
u
M t>
cd in
/4 Cu -1
^^
co co
i
o
cd CD
O
r/i J3
CD
m 1,111II. 11,1 t |f
!I!!I!M 'I 'ill ' I ' I ' I . I ' J ! EZ -
I'll I I l'l I I I'll II l'l 1 1' I ' I ' I ' I I ) I
ON
oo cn
en
-m 91- cn
oh

e
u
o
a
s
00
O0 u
/. ^S*\ ' 0O .' ' ' '
/' y^-'' O O B
'- * oo
Jl -.

* . t
0 ''".'
00

..

'.
. .

*
,',_

* *


S
5 tO
. " . -; 0
oo
/ oo
'. ' . ' . '
; -.'' 1*.
2 ' . . :

1 .
Oo 1 ' e-
. ".I".*..*. * <o?
*o * " .
o "' - 0" ' 1 1
SO ':. v '
-'Hi OO
0 n l l'l ' t . l f l ' l 'l
' . ' .

J I1 ' . . ' . ' . . ' .


: ;.' :" -' "oc
:i! I l| \\\ | M J t1 I ' . * , *

i'l

in
as cn
tu Z.I
m
212 IL60

IL55

10 m

333
Siltstone Sandstone Conglomerate Siltstone Sandstone Conglomerate

(a) (b)

Fig. 3.20 The Bulgo Sandstone cross-section in the West Cliff area, (a) borehole IL60, and (b)
borehole IL55.
Fig. 3.21 T h e Bulgo Sandstone in borehole IL60. The photo presents the middle part of the
formation consisting of fine- to medium-grained sandstone with prominent vertical and
subvertical joints with a length between 0.7 - 1.5 m .
JO
cd
U gw.a|
cd
u
a,
T3 w ^11
2> - > 2> fck
.ti 3 a a po

<u
u
^o 5 Sl00 3
E *
S3 u e 2 *
- J -2 i -s

8jj*s

21
3-3.2, 5 T3
a 3 .
2"1S
oo x> 3
5-s ""^
s'S g
o S u o
8 * g U

u 8 *<2u
cn
CN

on

di .; S*-i ^ ^J

3
<H so a
75 c 5
o Cj u <*
o .g -o
a .3 ^ c M
S S <u
3
8 o 0.5

. .3
11
Ji
U
cd Cj
US 'S H
co q a sg_ so
8c2 o -a -c
o =3 .-12

3 U'S 2lr2

"as
a M bgo al
'S
a i ^ g
o a- ,,
^ -o ,-
DC.S8STJ
5
'S 2 *
-a* 8^ Sf
v 5 i> oo ^ -o
fN
CN

00
FIGURES TO CHAPTER 4
agR co
<d .a
S 00
2.U
3 o
a 2

5" 2
i 2 ;
s:
9"J3
ag-b
UH

cd CO 13

a s.g
u
flj ... o
a
a CJ

O
<D

S a I
X) ^
cd cd fi

B
_^ O

8 CJ
a Q
co B -B a
U U ao o
| 3
CO CO
3 .j
.3 a
a. cd
2 3

00

U T3 cd
2
a 5 3 g
3
1>2
g co
.22 o
^ a.
o
ill?
3 oo i
So o 00

JfU cd
"-1 -o a

S 2 2 S-o
a cd -S O g
co Jg
o
I a a* a.
cd g

l.CS
.-S
1-1
xf >
co *- co 73.
if a oo a a.

o
CJ
ii
a
a
o <u
D O
3 -a B S
5^1
- S3 o

OO
c
c
8
fa
k t
j
<
00 o
B
c CO a
c o
0 c
co "1
-
4-*
>> B
.O S
T> pa
2.
c CO
Ba: B
u
<j
cd
I
8-3
o
aa
o cr
o, 3-
o
a
2$
3
a
1
OO
i XI
cj
a
_B O
OO a
a a
1
o o
1/1
a CM
c X
2 i)
Q.
'B
a O
2cd 8 .
o o
M
/c o Co
d\ 'a x
"3-
^f
M
oo
KL,

"8 2"2^

"2 "ft -*5.


't* li -
a-8 -8
jo-5 o>r>
8 --ad
a a <PM
_ O-fl
9"
1-1 s' - o.a .
**
<0 co o "S S3
iid X!
w5J *5- 5
S CJ <S-o
I .so 8
B rr co
a o o
az 00
rf

- 2 gas
1 . 2 * S 44
? .2
B
^
<H .a g . 2
oo"3 ^. oo
CJ ^. 0
^ 13
o -
i * 11 s
o
a* o.a c -s
<D .a =3 i- 2
- a ^ o
'a ) 6 E g

H
SI a -8
s 2 a
3 aSl s
oo
S3
cd

O 3 * .*
s oo^--' a
-a w <~N"O

2u
Ba
u cd oo ,
5
a a
5 2 3f2
5 * .
aj 8 e
a _ ~s ^
" SCob
ao
8 ia
CJ, - !H ^
a - ^
^8 o

3l"
u
00.2
a a o T>
&a 3d
3 A
53 =3 <t! o co
Il-a- a
o I-0* I
1 a-SZ V
w
-
"B, O
2 OB _
,c J3 a a o

OO

OPT) C

a | 3*3
g CO M _
o -o fl

erf* a co o

C-2 ^ g

:. -B 3
S 2 ,
" a
* 1
3 2
g ^ oo-g .a
**

1*33 -o

I laa-ou
CO
a g r.

^g
2 ! J <
a S^-s,0-
CJ o ^ op

.B cd cd
cd u "S- co sa
B p -1 a
i cj
"
>>cj *
a o c
oo??
00 d .a-S 9

op
S3
QUARTZ
QUARTZ QUARTZARENITE
QUARTZARENITE
SUBARKOSE SUBUTHARENITE
SUBUTHARENITE

ROCK FELDSPAR
FELDSPAR FRAGMENTS

(a) (b)

QUARTZ QUARTZ
QUARTZARENITE QUARTZARENITE

SUBARKOSE SUBUTHARENITE
SUBUTHARENITE

ROCK FELDSPAR ROCK


FELDSPAR FRAGMENTS
FRAGMENTS

(c) (d)

Fig 4.7 Classification of fresh,slightly and moderately weathered sandstones, (a) Coal Cliff
Sandstone, (b) Otford Sandstone Member, (c) Scarborough Sandstone, (d) Bulgo
Sandstone, (o) Fresh samples, (o) Slightly weathered samples, (A) Moderately weathered
samples
8-S 0

HI =1 a ,
\r> fl gB 00
-B
S.2f
co a, 3
S a
-a
a ^
p *!
a a a -i ex
. "o
D<ci-i o
B O 3
a 2 a"
CO a oa
oo o CJ -zi

3S 3 5 o
G o 2
cd T3 T3 O
o.a 2-S
cd B"
o u 2 <J-I
co cd
12
9 aa .3. d
co .3 -a _^
oo CO '

*1IJ5>-S
"* *2

2 =i CO *i
O cd

I o a
a. 5
I 9aJi-S CJ

a 73 9
3a co
MS a.a
CO XJ o
CJ cj
""" CO
co CJ -5
Si? cd .
cd a JO

p oa
oo 2 a
S3 .3 n a.

oo
S3
Fig 4.10 Moderately weathered Bulgo Sandstone (Table 4.3c sample No. M W B S s 7 )
between Coalcliff and Clifton. More intense weathering of the chlorite (Chi)
in the middle of the photo is accompanied by an increase in the amount of iron
oxide (10). Top - plane polarised light. Bottom - crossed polars.
20 -i

16

'B
U

1
' ' '''' M i i i i i i i i i i i i i i i i i i i i | i i i i i i i i i

0 1 2 3 4

Grade of the weathering

Fig 4.11 Decrease in calcite content with an increase in weathering for Narrabeen Group
sandstones, (1) fresh samples, (2) slightly weathered samples, (3) moderately
weathered samples.

10-i

X
o 6-
c
&o

T i i r~i i i n j r*T~r - 1 i n i i | i i i n i i i i | i n i i r n i r~

Grade of the weathering

Fig 4.12 Increase in iron oxide content with increase in weathering for Narrabeen Group
sandstones, (1) fresh samples, (2) slightly weathered samples, (3) moderately
weathered samples.
X

15

^
10

5-

i l i I i I T ' ' ' ' ' ' ' ' ' '
0 i I I I I I I M I | I i i ' ' ' ' ' ' l ' |-r
0 1 2 3 4

Grade of the weathering

Fig 4.13 Increase in matrix contend with increase in weathering for Narrabeen Group
sandstones, (1) fresh samples, (2) slightly weathered samples, (3) moderately
weathered samples.

12a

10 . F

O
u
CD

U F
SW
4^ \
4- \* MW

2-

SW MW
- -
D i i i n i n r 1 i i i i i i r~i i ] i n i i i i i r ] n ir~ii
7 12 17 22

% Matrix

Fig 4.14 Decrease in chert content with increase in matrix content for Coalcliff Sandstone
(CSs) and Bulgo Sandstone (BSs) accompanied by an increase in the weathering.
Fresh sample (F), slightly weathered sample ( S W ) and moderately weathered
samples ( M W ) .
CD

cr

a cn
4-1
ID
CM

m
<D
C3)
C<->Q

CN

> I-
O
^

CN
0>
0)
I

CO
o
CO X ~* r
o

CD ^ L
aj
o. r
P
5i
COo; o o o o
o o o o o
o CO CO CN
CJ O 3 C - M CO
y

Sample: B:\j2 10-29-1992


r 500-T7

U
t 400^ 1

3 '\\
K
i
300-
_j

]
i
I A
200-H '!
1
1
^h^l/^h\ H
/
\
!
i
] ^MJ,-. l v v -v
r *v>ji M i r> I i
100-1 ,Jr \ !
-I
o ] !I I l i i;i i i M i i i j i ii n i i i i i iI | ! II Ill I ! '

4 6 8 10 12 14

Sample: B:\j2-gf 10-29-1992


c 500
o
u
n
t 400-i
s
300- K

200-1
t> \

100- /
'~v^ v ^,^\y,^^ r ^ s ^ / ^ AJ ^^\^ , A-'j

0 - i i ; ' r -
! IIIII !iI IIII M II I ! I I iI I I !! I I iI !! Ii ! ! ! I I ! ! : M I
I I 14
10 12
4 6
Sample: B:\j2-ht 10-29-1992
C 500-
o
u
n
A
t 400-
s
300- \
\
;
200-
M

100-
w^Vf^^' 1 ^^
1 I I I I I M I ] !I I I I I I I I 1 I I I I I I I I | I [|1 II | || M !I I |I | i | ! I I I I I I I I I I
6 8
Degree^-Theta 12 14

Fig 4.16 X-ray diffractograms of sample J2 (talus sample from Jetty rock slumo) A
untreated, BB:
nntrp.ntp.ri. : glycolated, C: hpatoH
elvcnlsifpH P- heated tr>
to Anno
600 nC. v>
'
M = mix layer K = Kaolinite
FIGURES TO CHAPTER 5
Leaend

Qa Alluvium, gravel, sand, silt, clay


Qrt Colluvial/Soil and Talus
Rh Hawkesbury Sandstone -;Coalcliff
New port form/Bald Hill Claystone/Garie Tonstein^-
Bulgo Sandstone
Rnsp Stanwell Park Claystone
Rns Scarborough Sandstone
Wombarra Shale ur Fault
Coal Cliff Sandstone
Illawarra Coal Measures
ty Fault

Fault

Fig 5.1 The location of major faults in the slip area (modified after B o w m a n , 1974).
"*+n^e' 5
. SEES

ft>3 Xv
Fig 5.2 Small fault in the Bulgo Sandstone above
Lawrence Hargrave Drive between Clifton and

Coalcliff (see Fig. 5.1 for abbreviations).

,-r~- (Mil t Fig 5.3 Small fault in the Stanwell Park Claystone
above Lawrence Hargrave Drive between Coalcliff and
Clifton (see Fig. 5.1 for abbreviations).

Fig 5.4 T h e Harbour Fault to the south of Coalcliff Beach. T h e fault has cut the
rock platform with a readily disdnguished lineament (sec Fig 5 1 for
abbreviations).
c- < ^ The Jettv Fault (left) cutting the Otford Sandstone M e m b e r ( O S M ) on the

about 70P to the north and strife is east-wes The fau t J^ * J^, in the
groundwater circulation under the r o a d ^ T h ^ J ^ y ^ faPult p la e

abbreviations).

Fie 5 6 The Clifton Fault to the north of Clifton. The strike of fault is east-west
" with a dip that is nearly vertical. It is marked by a prominent, straight
creek This fault has caused an increase in the water flow and is directly
related to the Moronga landslide (the left of the photo; (see Fig. 5.1 for
abbreviations).
N Scarborou

Scarborough J?
Fault "

c
ft

1000 m

Fig 5.7 The Scarborough Fault has a throw down to the north. M a x i m u m dip is
60 toward the north. The photo shows the northern split of the easterly
directed fault to the north of Wombarra Beach. This fault appears to be
related to the Wombarra slump (see Fig. 5.1 for abbreviations).
N \ /X^^Stanwell Park

\ _
V_^7^^
^^A ^A^J
^^Coal c1i ff

T vA.
5 ,T Harbour Fault
1
/\8 AA1-2
9 Jio Jetty Fault

_>13
C].ifton Fault ~t/~~~^~ 1 4
"/Scarborough
\

\
Scarborough\
Fault ' \ '

\ AAA
^AAf
c
f\ / .0
C
9 ! \ /
y Wombarra
/ <D f -c
<-> /
/c /
/ ru

/ F Coledale
| 1000 m ,

/ iff
Fig 5.8 Location m a p of sampling sites for joint measurements.
Fig 5.9 Rose diagram of joint orientations for the Bulgo Sandstone (113 reading,
from sites 1, 2 and 3, Fig. 5.8) .

Fig 5.10 Rose diagram of joint orientations for the Stanwell Park Clavstone flOO
readings, from sites 4 and 5, Fig. 5.8).
Fig 5.11 Rose diagram of joint orientations for the Scarborough Sandstone (133
readings, from sites 6, 7, 8 and 9, Fig. 5.8).

Fig 5.12 Rose diagram of joint orientations for the Wombarra Shale (110 readings,
from sites 9 and 10, Fig . 5.8).
- /

r-

i-Sf
T^fc
"Jf tp :::: 3
-I
\ \

-- /

Fig 5.13 Rose diagram of joint orientations for the Coal Cliff Sandstone in cliff
exposures showing dominance of joints parallel to the cliffline (90 readings,
from sites 10, 11 and 14, Fig. 5.8).

Fig 5.14 Rose diagram of joint orientations for the Coal Cliff Sandstone measured
on the shore platform (114 readings from sites 12 and 13, Fig. 5.8).
Fig 5.15 Total joint directions for Narrabeen Group in the study area (from 14 sites
between Clifton and Coalcliff area, Fig. 5.8).

Fig 5.16 The joint strike maxima at 015 for Wombarra Shale has a great effect on
slope stability between Clifton and Coalcliff. also for Coal Cliff Sandstone
and Scarborough Sandstone as seen in photo.
/C XC / s / v
a) Jointed rock mass, no pronounced X >^\\
effects of previous shearing displacement

b) Similar rock mass, fault present;/ J

Fig 5.17 Significance of faults in slope stability problems (after Patton and Deere,
1971).

Unaffected (country)
SIGNIFICANT STRENGTH &
PERMEABILITY CHARACTRISTICS rock (e)
Fractured rock (d)
Slickensided, striated
surface (c)
Average permeability Fault Gouge (b)
& strength of country
rock.
Fault breccia (a)
(with gouge)
\ Fault gouge(b)
- Slicken, str. surface (c)
Fractured rock (d)

High permeability
zone
Low permeability Unaffected (country)
low strength zone rock (e)
Low to high perm.
low to moderate
strength zone
Low permeability
low strength zone
High permeability
zone

Fig 5.18 Typical cross-section of a composite fault (after Patton and Deere, 1971).
SB o
o
4-1
c
o
Xi
U
c
8
if
>
=i XJ
'% g
00 d
e Q
n
cd 1
? /-N
cd CO
<u as
S a
>> c
X> i2
I cd

0 S1xi
J
T3
C
a
IH

9 S
i

so
J3
00
00
o
2* 18
Jo2o.>
xi
_2
sSB
73
t-l
Sj -a
3 c3 Uo
o
<N
w-i
01

iS

Ui
c
cd <0
is
<D
sso 1
J3
00 Q
'2
cd
w
00
'$ ^-N
CO
CJ DC
CD
c
g o
E o 2>
o ed x>
ej
o
*CO
.2
C
1- s
a .2
- 0 ii

a 'S vS
'B
CO 4-t
3 eb
t-l

E
CO
g
N'
CO OO CO
I8
ft
B
3 r^4
cd
c o "o
s CO a
CO o
<D
a o u
T3
*3 i
ca P S
c 1=
o CMQ
v. ^
u 2o.O
s
o\
T

m
00

E
i t-l aJ
C4_i rt3

^ fl
H <D O

-DC
.a 8 |
.S8S
o .s
73
S8
OOx,
J3
C e o
SO
gs*
3
a
o^ &
cd oo
I?
O fl o
S8.t
*
3 jy *r* c
/>?P
K
a. <D J
o 8 .
BUS
.S3 8 J"2
00 3 ^
1 o .Ufc

CN

00
E

xfH o
CO fl B

.3 ~.
> . 00

- jj 8
a)
3
x! " S3
U.U3
00 h
cd o
.5 e U
2'"
"O fl -o

aIa
s P
CI, -^

co

b a
x^ CO
CO >
cd o
5
a * c ^.
;s.s Q ao
3 C
-2
CO CO
2 >
O T3
<N &
IT) DC-S
00

E
(W-E)
Zone of extension:
Loosened rock
Vertical to subvertical tension joints
Bulgo Sandstone Bedding planes

\
Stanwell Park Claystone
\
Scarborough Sandstone
\
Wombarra Shale Major bottom deformation
\
Coal Cliff Sandstone
(open bedding planes)
Bulli Coal Illawarra Coal Measures

(a)

Tension joints closer spaced


Bulgo Sandstone
Diagonal to curved
Stanwell Park Claystone shear joints

Vertical to subvertical
Scarborough Sandstone
tension joints
Wombarra Shale
Coal Cliff Sandstone Major bottom deformation
Bulli Coal (open bedding planes)

(b)

Tensile fracture pattern due t;


Bulgo Sandstone destressing of rock

Stanwell Park Claystone

Scarborough Sandstone
Slump slide
Wombarra Shale
"Major bottom deformation
Coal Cliff Sandstone
(open bedding planes)
Bulli Coal

(C)

Fig 5.23 Schematic escarpment cross-section (modified after Ferguson and Hamel,
1981).
Zone of opened joints JI'I'I
(W-E)
-H-''/.// Hawkesbury Sandstone

Illawarra Coal Measures

Major bottom deformation

<l

Fig 5.24 Horizontal stress plus vertical load removal causes arching and buckling of
beds in the escarpment bottom.

Fig 5.25 Sandstone cliffs along the sea are vertically jointed and break leaving
vertical faces. Rockfalls occur in these places because the toe of the slope
is eroded by the sea; relaxation of the material above produces toppling
along joint faces (see Fig. 5.1 for abbreviations).
FIGURES TO CHAPTER 6
r-
o\
r-H

a
cd
I-
XL -a
CD
<U
*-/ c

l_ 2
LU
TO CJ 1-
<
SJ c 2
Q CO
c
cn cd
_ ~-
CD u .
Z. LU
X. V .c n CL C
u T3 4-1 o o O
CO CL
L. l/l CJ
ex. _l
tD CO
1_
CO T3
CJ > cn J Z u
4-J CJ
c < ^
C4-1
TO .l
2 L. /l o
2 cd
T3 ro Q- O
C CJ OJ C/3
3 4-1 L. > -1
0 CD u h- Ll_ Q.
L. 2 c
cn


O
au
0
h-
H _l -a
3 LL' c
CQ 1 3
_l p
S-i
LU <
ex C4-I
Z3 <
1- D- O
< c
ex Q 0
LU LU Cfl
H S c
1 Z) cd
_I LU CO Cu
U CO
_l < 6
0
< H LU
u u 2 1
O Cfl
u
<
z ce _l<
X 0_ LL. h- ^H
CJ cx ex vd
h- 2
LU LU LU
i-H
O
LU Q l-H
< < PH
tD 2 3: 2
ZD QC3
2 O 2 z:
LL.
0 0
z 2: ex ex.
0 0 tD CD
2: Q
S:_J
O_)LUCO 'cd
( ^^^
CO
_^ C3 TJ
03
LU cu -^
a s ^
LU (L)
a a p

ressu
Of si
t,

_i
CO

u_ tu o,
o ccd ao
LU
_J E o-
h- 4> """

z.
<
>* c/3

he i: crea
slide and
r-
.. c
2

i*1-"N
+->
JD <U rn
C 5CD
O 3
CO ^ cd
. u 2
pe b
J3 llust
TO
4-1
L.
CJ o -*
4-1
TO c
11 ,
2
3 C ^.{S
able landsli e debr
and Hendr n, 197
of groundw ter flo

CT 3
o
L.
cn
II cd cd "O
h-
CJ

c c <u
o
C/3
rl cd <U
4_> tn

LU u/ w H
Q & us
1
CO
E a-
O <+-i dJ
W V-' 41
LU
CC
O CN
LL
LU
P3 *00
d
cd E

3
CT
CJ
Potential slide

Sea level

\ / */ '
\l
Fluid pressure
distributions

6.3 Possible effects of high fluid pressures at the base of an escarpment in the
groundwater discharge area.

a) Fault acting as a
groundwater barrier
due to fault gauge

b) Fault acting as a
groundwater conduit
through the fault
breccia

c) Fault breccia acting


as a subsurface drain

6.4 Different effects of faults on groundwater conditions (after Patton and Deere, 1971).
Fig 6.5 Water flow within the rockmass is concentrated along discontinuties at the bases of
the sandstone units, for example between the Wombarra Shale (below and behind the
shotcrete and concrete retaining wall) and Scarborough Sandstone (top). This increases
the rate of weathering of the Wombarra Shale, causes fretting and weathering of the
sandstone and leads to toppling and rockfalls. The weathered shale has been faced
with concrete. Lawrence Hargrave Drive between Clifton and Coalcliff.
(a)

Perched
water
table
Main -
water
table

(b)

40 Scarp crack

Piping & collapse of


talus along scarp crack
Road

(c)

6.6 (a,b) Coledale Station on the Illawarra area acts as a local catchment to replenish the
groundwater table within the escarpment slope, (c) A deep seated failure has been
active for some time at Coledale, this was exacerbated by the heavy rainfall in April
1988.
I w y ' February ' March ' Apn7 ' May ' Juno Ju* August Septet October November December
Month
1990 1092 1994
1985 1986 1987 1988

m
Fig 6.7 Rainfall data for the Illawarra area. 1988 involved both high intensity and long
duration rainfall events. 1991 was mainly dry with only one major rainfall event in
June, Which resulted in many new debris flows. (Wollongong University Station).
600

500 -

400

E
E,
= 300
c
'5
m 200

100

January February March April May June July August September October November December
Month
1985 1988 1990

Fig 6.8 Rainfall data for the Illawarra area. 1988 involved both high intensity and long
duration rainfall events. 1991 was mainly dry with only one major rainfall event in
June, Which resulted in many new debris flows. (Clifton Station).

y = 11.088+ 0.85738x R A 2 = 0.881


e
0

200 400 600 800

Rainfall ( m m ) from Wollongong University Station

A positive correlation between rainfall in Clifton and Wollongong areas during the
period 1985-1991.
1! t&''~yr:^*'&$i%-i IIP?
xr ^"}?*'' |flifcjj'<"' i W^'l^'''

^ffiaSaflflfi
- ^

4t' '

N ,

Colluvium

.__-_^~ Weathered Wombarra Shale

Coal Cliff Sandstone

Illawarra Coal Measures

Fis 6 10 Water flow within the soil mass is concentrated at the base of colluvium (between
' Scarborough and Clifton). Just above the Coal Cliff Sandstone Wombarra Shale is
present the water should escape into the joints and sandstone itself, this m a y be due
to the slight northwest tilt on the sandstone. The base of colluvium (top) lies on the
Wombarra Shale and water drains from the base of colluvium. Slope instability is
aided by the high water table.
"TS'V S3bi3H
S 3 5 o S 8 -* 3 S

3
O
CD 00 in
M
P -S CD
O OO o oo O
a
C i- a cd 2
C/3 2 oo cd c/3
r
cd 00 " *
<i j 4-4 3 3
!i
X T3 <D
11 CD ^3 CD 3
a
s i-l c O
cd Ca
TJ on
c O
OJ O
<
=u
5 T3
73
o
*s <D r j
D >
X) cd 3
<
C U O
CL) cd G c/j cd
J3 E
<U S-i i cd O rv cd
* * CX\
Q, 00
.3 cd
o
CD
.a
CD t_w 3 E^
3 '-3 P 3 32
cd
C/3 .3 cd
00 p <5 J2
s o cd C4-I *
o ft)
_ 00 3 5H 3
*-l

cd . o w 33
<u CD CD o ^3 cd
*"* cd _e cd o cd <
<D ftJ
J3

a cd
S :? oo
O O w
cd > -3_ C4-I < 4 ^ 00
CO
C/3
-!
c
o
-a u-s
o o -a
O s 35
LL 00
CD 1 *" ~ 2CD
3
O
00
00 cd
s 3 "
'o ftJ
-^
O "ccl ^ a s > <-' ;- CN
CD 3 0) 00
CD cd
cd oo E !t3 ON
o i- <->
|-H O cd a n1^ 3 '
CU -a 3
M
2 4J
C4_| "T3
a z CT r
cd 00
DO
f-H * Id
cd ^

00

c I
E
1 r\l?r
c 5
0
O UJ
fi\ ZJ

J ''X
/ i\
i\ o
o
LD

c
^!N
0
x;
f
til \
0 "C a*
V> V * 1

f; "5
11
3 0 3K
m :i .
O Or
a. -D
.1
~ 0 0
e
0 < 1 *- * 2
C =
0 0
' :l O
(/I

g 3 A B S
TTT 8 3 S ^"i"
\
1950
B L956
1990

1963
1988 a n.
\ 1939
91

WL
Oyj

& D MONTHLY 9 3 &NIHLY

i i i i i i i i i i i i i i i
1.0 10.0 100.0
Recurrence interval (years)

6.12 Rainfall vs recurrence interval (years) for Woonona/Coledale Station during the
1930-1990, (after Longmac Associates, 1991).
Fig 6.13 Isoseismal m a p of the 9 March 1973 Picton earthquake. Small figure next to open
circle indicates intensity different from zone designation (after Denham, 1979).
DATE: Zl MAY 1961 A EPICENTRE
TIME: 21: 40:020 UT IV ZONE INTENSITY DESIGNATION (MM)
MAGNITUDE: 5 8 M B , 5 6 M L
HYPOCENTRE: 34 55S 150 50E
DEPTH: 19 km

Fig 6.14 Isoseismal m a p of the 21 M a y 1961 Robertson - Bowral earthquake. Isopleths show
zone intensity limits (after Denham, 1979).
DATE 28 December 1989
I- TIME 23:26:58 1.5 s UT
MAGNITUDE :5.6 ML
EPICENTRE : 32. 95 S. 151.61
DEPTH: 11.5 1.0 km j

A Epicentre
IV Zone Intensity Designation
3 Earthquake Felt (MM)
30
o Earthquake Not Felt

100 km
I

Port Macquarie

32

34

o
2-3

Wagga Wagga O

Batemans Bay
s NSW 36^
"'^Albury \
VIC
.0
146 148 152 154
_L
Fig 6.15 Isoseismal m a p of the 28 December 1989 Newcastle earthquake (after M c C u e et al.,
1990).
FIGURES TO CHAPTER 7
^
-
Coai S95
? CD v ^ " \ ^ ^ SW
C
'
2 90-
* GO
B 8.
i"
I 85 -
r x. ^ \
~*^^
3
Sandstone a-
15 \. MW
n 20- 2
(0
80 -
T3
O
^ ^ - * ffi
1 2 3 4 2 J "
No. or Slaking cycles " 75- A
70 No. of slaking cycles

(a) (b)

Fig 7.1 Effect of number of cycles on slake durability (a) for upper Illawarra Coal Measures
(coal and interbedded highly weathered sandstone), (b) for Coal Cliff Sandstone
(interbedded claystone, sw = slightly weathered; m w = moderately weathered).

2 3
No. of slaking cycles 2 3
No. of slaking cycles

(a) (b)

Fig 7.2 Effect of number of cycles on slake durability (a) for Coal Cliff Sandstone (sw =
slighdy weathered; m w = moderately weathered), (b) for Wombarra Shale weathered
samples from different locations, (top)from Clifton area beside Jetty Fault and (bottom)
from south of Wombarra Station.
2 3 2 3
No. of slaking cycles No. of slaking cycles

(a) (b)

Fig 7.3 Effect of number of cycles on slake durability for (a) Wombarra Shale samples, (b)
for Scarborough Sandstone (sw = slightly weathered; m w = moderately weathered).

100

2 3 2 3
No. of slaking cycles No. of slaking cycles

(b)
(a)

Fig 7.4 Effect of number of cycles on slake durability (a) for Scarborough Sandstone
(interbedded claystone), (b) for Stanwell Park Claystone ( m w = moderately weathered;
h w = highly weathered).
t=r _

SW -^ITrrrE^t -
"k .
MW
92 - - -

2-
- ^^.

^"\

J i 1 , i
2 3 2 3
No. of slaking cycles No. of slaking cycles

(a) (b)

Fig 7.5 Effect of number of cycles on slake durability (a) for Stanwell Park Claystone samples
from different locations, between Clifton and Coalcliff beside Harbour Fault (open
circle) and from northern Coalcliff Station (filled square), (b) for Bulgo Sandstone (sw
= slightly weathered; m w = moderately weathered).

No. of slaking cycles

Fig 7.6 Effect of number of cycles on slake durability for interbedded claystone in the Bulgo
Sandstone (weathered samples).
OSM SSs CSs SSs SPC
Name of formation N a m e of formation

(a) (b)

Fig 7.7 Slake durability index (a) for Narrabeen Group fresh samples (West Cliff, boreholes
IL55 and 1164), (b) for Narrabeen Group weathered samples (between Clifton and
Coalcliff).

4.28 5.05 775


Water content ( W % ) Water content ( W % )

(a) (b)

Fig 7.8 Water content versus two cycles of slake durability (a) for Stanwell Park Claystone
( m w = moderately weathered; h w = highly weathered), (b) for Wombarra Shale
weathered samples. (Samples were collected one week after rainfall).
a,
R
"a.

CO

(4-1

S.
0
t
a,

S
a
o

tm.t^s< 1
AAA, CO j8


oo o
a a
8 Oa
<H
cd
h) u_.
co o
3 cK

oo
IE

E
cd

a)
j:

a
0)

sa)
e
o
CO
X)
c
cd
CO
TJ
CI
a
Ou
O
X)
CI)
c co
Pf H
1)

i<
J4
a.
>,E
o
C-
u
4-1

Ee
cd
an . >
o a
Ct-l

u
c c2
oa 8
CO
3 *3
>-M O

o\
r^
on
li
O
o
K
<H 2
JJ o
s ^
CO L*
CO >

{J CO

8-s
.o -

53 J2

II
g^
2 "
3 c3
..
XJ Cu
o
co w

<-> <u
SB'S

.^I^I^I^I^H r-
oo
E

00

:=?S

5
CO \ /
ed xj
U cd

a
icS
2
(5 2
o --'
^ co
3 "S"
i
co
Ui

o o ,
a o t.
3-8
S^XJ

111
P3 S X J
oo
U
o
XJ _C0
c 3
,c CP
U

o ca
OJ 3
XJ
a; c XJ
n E U
E E cd
CO
u
CO
b co
J=
c/:
"

c EF
u
dP
_ J= c
oo
- S 5)
x:
c i*
B (3) ssof iifSra/w 0/ 3
CO
cd X J (3) ssoj ; L I 3 ! 3 ^ p a
oo CO
c CO
u lcd CO X 3J
on /J cd cd
CO C/1
jr
Oil u
D.
3 * XJ
XJ n
.2 c XJ Cd C
: l 1 i 1
c
c O
=Z i l l ,
c 00
c c
_3 OJ
! i 1 1 1 1 I c OJ
o u 3 91 f-i X:
CO CO
n j

>%t
>, U
= E
XJ E
MUM/ CO
>,
XI
*u-'
x:
X)
OJ
z
u -
3 C
cd OJ
CJ JS
! ! ! ! ! W. X!
CO X)
3
cd
U

E
a
y;) 3
3
cd
o X)s
,*"'
(3) ssoj iq3[3M CO * i < ! ! ! A ioi CO
0)
CO
CO O 3
o J
oon
a I | V O

_ c
00 CO
3J
3

XI
- j

x; XCOJ
CO
u
O

1 / /!
| -s
oo
1 ' 1 i
E l
r- oo
i j E
oo
i
E
r j
i
|
i
1 [i
i i . i . i
C3) ssoj aq3ia^\
(3) ssoj jijaiayvj
L > O.50

Equivaienr core

0.3 W < 0 < W

(a)
(b)

L > 0.5 0

Equivaier

Section through
loaainq points
<0 < W
W
(C)

Fig 7.16 Specimen shape requirements for: (a) the diametral test; (b) the axial test; and (c) the
irregular lump test (after I S R M , 1985).

(a)

(b)

(c)
(d)

Fig 7.17 Typical modes of failure for valid and invalid tests (a) valid diametral tests; (b) valid
axial tests; (c) invalid diametral test; and (d) invalid axial tests (after I S R M , 1985).
8 Y 5 -/

4-
6
1
3 -

<\ I S
1 B:
^1 ^B^

1 II B H H Hi
is
CO 2-

2 -1 I I Hi |
1 -

1 !
^^
II 1 1 li
0
xjifl HI
0 -
CSs WSh OSM SSs SPC BSs CSs WSh OSM SSs 1 SPCB P 9BSs
Wis?

Formation Formation

(a) (b)

Fig 7.18 (a) Axial point load strength results, and (b) diametral point load strength results for
Narrabeen Group (fresh core samples).

00 -/ ^ |

80 -

60 -

c
40 -
1 B
20 -
^H:;'
B
B
/jiMBl IBH

CO CO CO O
0-
ISxBB Bf> IB
CO
j? CO CO CO
.CSs

CO CO CO CO CO CO CO 2
CSs

0- CO CO CO CO o CO CO
Oo g CO CO
o
5 co m m g CO CO CO Q.
tg
5
1 2
%
g
CO
g
5
g
2
g
CO
g
5 I 1 n
g g CO
m m
CO
Formation ^ FORMATION 2

(a) (b)

Fia 7 19 (a) Point load strength results for Narrabeen Group (weathered irregular samples), (b)
' Uniaxial compressive strength for Narrabeen Group (weathered irregular samples).
CSs WSh OSM SSs SPC BSS

N a m e of formation

Fig 7.20 Anisotropy for Narrabeen Group point load strength (fresh core samples).
120

100

VI
O

Fresh S. weathered M.weathered

Weathering

(a)

Fig 7 21 (a) Uniaxial compressive strength (UCS) versus weathering for fresh, slightly and
moderately weathered Coal Cliff Sandstone, (b) Uniaxial compressive strength versus
weathering for fresh and slightly weathered Otford Sandstone M e m b e r and moderately
weathered Otford Conglomerate. Numbers in brackets are the slake durability index.

Fresh M.weathered

Weathering

(a)

Fig. 7.22 (a) Uniaxial compressive strength (UCS) versus weathering for fresh and moderatel
Stanwell Park Claystone. (b) Uniaxial compressive strength versus weathering for fresh
and moderately weathered Wombarra Shale. Numbers in brackets are the slake
durability index.
M.weathered Fresh S. weathered M.weathered
Fresh S.weathered

Weathering
Weathering

(a)
(b)
Fig. 7.23 (a) Uniaxial compressive strength versus weathering for fresh, slightly and moderately
weathered Bulgo Sandstone, (b) Uniaxial compressive strength versus weathering for
fresh, slightly and moderately weathered Scarborough Sandstone. Numbers in brackets
are the slake durability index.

200 - /

"c? HH HB#
0.
H. <ool
100 1
C/3
U
3

Hi
111 mm
0-
CSs WSh fl
OSs fl?
91SSsiH> SPC BSs
CSs WSh
OSs SSs SPC BSs
Name of formation

N a m e of formation
(a) (b)

Fig. 7.24 (a) Uniaxial compressive strength from axial point load strength, (b) Uniaxial
samplesT1^ ^"^ ^ ^^ Pint l0ad Strength f r Narrabeen Grou
P (fsh
_12d

i
a? 100

SSs
CSs WSh OSM SSa SPC BSs N a m e of formation
N a m e of formation

(b)
(a)
Fig. 7 . 2 5 Comparing uniaxial compressive strength (UCS) and slake durability (SDI) (a) for
Narrabeen Group fresh core samples (West Cliff, borehole 11155), (b) for Narrabeen
Group moderately weathered samples (between Clifton and Coalcliff).

60

40-

o
3 20-

I i I i I i r i i i i ii i i ii i i i r i i

66 68 70 72 74 76 78 80 82 84 86 88 90
SDI (% d 2 )

Relati
Fig. 7 . 2 6 nship between uniaxial compressive strength (UCS) and slake durability (SDI)
for Wombarra Shale moderately weathered samples.
30

O 20-
CL

V)

0 10 20 30 40 50 60 70 80
SDI {% 62)
Fig. 7.27 Relationship between uniaxial compressive strength (UCS) and slake durability (SDI)
for Stanwell Park Claystone weathered samples.
132.00 z\ 85.00

80.00
1 28.00

75.00
124.00 D
a Q_
CL 70.00 i

120.00 -;
00 00 65.00 i
o 116.00 - O
D
60.00

1 1 2.00 i
55.00

1 0 8 . 0 0 1111111 n1111 n n 11111111 n 1111 i-rrrr 11 m 11111 M 111i1111111 11 I|i 111 50.00 "1111 M II 1111 M 111111111 i M 1111 M 11111111 M 11111111
25.00 26.00 27.00 28.00 29.00 30.00 31.00 21.00 22.00 23.00 24.00 25.00 26.00
% Quartz % Quartz
(a) (b)
Fie 7 28 Relationship between uniaxial compressive strength ( U C S ) and percentage of quartz
(a) for fresh samples, and (b) slightly weathered samples from the Coal Clitt
Sandstone.
1 70.00 180.00 n

165.00 - 170.00

160.00 z 160.00
D D
CL

1 55.00 150.00
00 00
O O
D 1 50.00 D 140.00

1 45.00 130.00

1 4 0 . 0 0 - i i i i 1 1 1 1 1 1 11 i i i 11 i i i i i i i i i i i i i 11 i i i i i ' i i ''' M ' 1 20.00 11 in 11 II in 11 n i n i ui i in m i 1111 n i II n|i i II in 11 II i II n n 11


35.00 +0.00 45.00 50.00 55.00 25.00 30.00 35.00 40.00 45.00 50.00 55.00 60.00
% Quartz % Quartz
(a) (b)

Fig. 7.29 Relationship between uniaxial compressive strength (UCS) and percentage of quartz
(a) for Scarborough Sandstone (fresh samples), and (b) for Bulgo Sandstone (fresh
samples).
180.00 80.00

160.00 -
70.00

140.00
D D
CL CL 60.00

120.00
00 00
O O 50.00
Z) 100.00 - z>

40.00
80.00

60.00 i i i i11iiii11i I I 11 i i i 11 i i i 1 1 i i i 11 i i i i i i i i i i i i i i i i i i 30.00 111111 II 11111111111111111111111 M M 1111111111111111111111


10.00 20.00 30.00 40.00 50.00 60.00 15.00 20.00 25.00 30.00 35.00 40.00
% Quartz % Quartz

(a) (b)

Fig. 7.30 Relationship between uniaxial compressive strength (UCS) and percentage of quartz
(a) for fresh samples, and (b) for slightly weathered samples from the Narrabeen
Group Sandstone.

35.00 -j

30.00

D
CL 25.00 :

00
O 20.00 :
z>

15.00

1 0 . 0 0 | i 11 i 1 1 1 1 i i 11 i i i 11 i i i i i i i i i i i i | i i i i i i i i i i i i i
12.00 16.00 20.00 24.00 28.00 32.00
% Quartz

Fig. 7.31 Relationship between uniaxial compressive strength and percentage of quartz for
Narrabeen Group sandstones (moderately weathered samples).
8.00 -n 4.00 -i

7.00 - 3.50 -_
X
(1)
d
a
c 6.00 3.00
X3
+J bo
til
C
c b.ao {5 2.50
eW
+> w
T3 cd
cd 4.00 2.00
n
+->
c o
o 3.00 OH
1.50 T
DH

2.00 - 1 1 1 i i 11 i i | i i i i i 11 i 11 11 i i i 11 i i i 11 i i 11 i i i i 1.00 11II 1111111111111111111111111111111111111111111111111


10.00 20.00 30.00 40.00 50.00 60.00 15.00 20.00 25.00 30.00 35.00 40.00
% Quartz % Quartz
(a)
(b)

Fig. 7.32 Relationship between point load strength index and percentage of quartz for (a) fresh
samples, and (b) for slightly weathered samples from the Narrabeen Group Sandstone.

1.60

X 140

tJ, 1 -20
bfl

c
K
w
T3 1.00
cd
OO 0.80

a,

0.60. 1 ' " ' " " ' I ' 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1, 111,


12.00 16.00 20.00 24.00 28.00 32 00
% Quartz

Fig. 7.33 Relationship between point load strength index and percentage of quartz for Narrabeen
Group sandstones (moderately weathered samples).
FIGURES TO CHAPTER 8
g
u
E
c
c
>
cu

Sea level

(a)

40 Scarp crack
Piping & collapse of
E talus along scarp crack
c
c
o
3

Sandstone

(b)

Fig 8.1 Cross-section of talus slopes in the Illawarra area:


(a) interbedded strong and weak rock (lower Narrabeen Group); (b) thick interval
of weak rock (Illawarra Coal Measures).
(after Hamel, 1980)
W-E
Tension crack
(a)

Water table
Talus

bedding plane Vertical fracture


(potential slip -
Seepage
surface)
Weathered bedrock'

(b)
talus-rock interface
(potential slip surface)

(c)
-Main scarp

^ ^
." . . .* ',
''.' . ' . . ' ' . . '

^^V*
* * * . .
. ,

, -
'''.' ' "
- .
. ' . '

'*."' * \ \.

Fig. 8.2 Engineering geological failure model for talus slope instability along the
northern niawarra coastline, (a) Initiation of tension cracks in talus and their
relation to vertical fractures in bedrock. Slip surfaces partly or totally pass
through the bedrock, (b) Slip surface at the contact between bedrock and
talus. A wedge of talus is developed in front of a drop in bedrock, which
represents a 015-020 joint (Fig. 5.16). The weight of the wedge causes a
downhill movement after heavy rain when the water table rises, reduces the
shear strength of the talus material and the wedge becomes detached from
bedrock at its back, (c) The tabular vertical gap formed between the bedrock
and the soil stretches the soil above, and forms a NNE-oriented fracture
which marks the crest of a future slide (Fig. 8.39).
Sample: B:\GHOBADI\ch2 07-27-1994
C 2000-

1600-

Kaolinite

1200-

800-
Swelling clays
\

400- i0H^r "^Atfrw^m^^ *%4t

1111111111 n n 11111111 n 1111111111111 ri p 1111111111111111111 rr 11 n 11 r m M M i M 111111111111111111111111111111 n 11111111111111 ITI 111111111111111111 in 1111 n i ITI 1111111111111111
2 4 6 8 10 12 14 16 18 20
Degrees 2-Theta

(a)

Sample: B:\GHOBADHch1gly * 07-27-1994


2000-

1600-

1200-

Kaolinite
800-

400- WwW%M^
muMrtfV
w^w^V^
~| 111 n 1111111111111111111 111111 u 111111111111111111111111111111111 -i 111 111111111111111 111111111111111111111 n 11111 m Trm 1111 m 111111 m 11111111 n 11
2 4 6 8 10 12 14 16 18 20
Degrees 2-Theta

(b)

Fig. 8.3 T h e X-ray diffraction trace of clay samples from the Clifton earth slump (a) before
glycolation and (b) after glycolation.
100n

>
H
81.33 KPa
80-
U
ft
67.7 KPa
60- 54.1 KPa
co
40.48 KPa
CM
26.85 KPa
>
40-
H
T)
G
H 20-
R3
0
0 f I I IIIIIIIIII III II IIIII I | I I I I I I I I I | I I I I I H I I | I I I I I II I I |
0 200 400 600 800 1000 1200
Horizontal displacement x 10" mm

26.85 KPa

111111111111111111111111111111111111111111111111111111111111
) 200 400 600 800 1000 1200
3
Horizontal displacement x 10"
mm

Fig. 8.4 Plot of load and vertical displacement vs. horizontal displacement for samples from
the head of Clifton earth slump (test 2, five repeated measurements, see appendix).
160

>
H ~ 81.33 KPa
120
CD
a
-__^ 67.7 KPa
00 80- . . 54.1 KPa
* 40.48 KPa
>
H 26.85 KPa
40-
C
i-i

cd
0
i i 11 i i i i 11 i i 11 i i 11 i i i i i i i l i i i i i i i i i i
0 400 800 1200 1600
Horizontal displacement x 10 -3 mm

50-i
26.85 KPa

40.48 KPa

c -50-
CD
g
u ^ * 54 . 1 KPa
rd -100-
.i

a ' 67.7 KPa


CO
H
(0
T3 150- + *
81.33 KPa
CJ
H
+J
l-i
OJ -200 i 11 11 11 i i 11 11 11 i i 11 i i 11 i i 11 i i 11 i i 11 i i 11 i

> i 400 800 1200 1600


Horizontal displacement x 10~3 mm

Fig. 8.5 Plot of load and vertical displacement vs. horizontal displacement for samples from
thecrownof Clifton earth slump (test 3, five repeated measurements, see appendix)
160
81.33 KPa
>
H
~ 67.7 KPa
PI 120-
CJJ
ft
54.1 KPa

t> 40.48 KPa


CO

26.85 KPa

C
H

cfl
0 0 fiii 111 n 1111111111111111 " 11111111111111 11 i|i II 111 i I'M
0 200 400 600 800 1000 1200
3
Horizontal displacement x 10" mm

Fig. 8.6 Plot of load vs. horizontal displacement for samples from the toe of Clifton earth
slump (test 1, five repeated measurement, see appendix).

90

80 r

to
ft: 70-
ix,
"'
(0 60 i
to
CD
u
CO 50 -. Test 2
u
(0
CD 40-
c
in
9.
30-

2 0 "| III III I ll| 11IIII III |ll II III ll| II III llll| III III II l| I III III II |ll III 11 ll| III II ll|||
10 20 30 40 50 60 70 80 90
Normal Stress (KPa)

Fig. 8.7 Plot of shear stress vs. normal stress for Clifton earth slump.
200 -i

81.33 KPa
67.7 KPa
54.1 KPa

- 40.48 KPa

. 26.85 KPa

0 f 111 11111 11 111 11111 1111111111 11 111 11 11111 111 11111 11


0 200 400 600 800 1000
3
Horizontal displacement x 10" mm

(a)

150-1
81.33 KPa
.67.7 KPa
54.1 KPa
40.48 KPa

26.85 KPa

o 0 f 11 11 111 11111 11 11111 1111111 M i 111 11 11111 11111 111111


0 200 400 600 800 1000

Horizontal displacement x 10"3 mm

(b)

Fig. 8.8 Plot of load vs. horizontal displacement for samples from Moronga Park slump-earth
flow, (a) test 1 from top of slump and (b) test 2 from head of slump (five repeated
measurement for each test, see appendix).
160-i

, , t 81.33 KPa

67.7 KPa
**S=Z=S
54.1 KPa
- 40.48 KPa

26.85 KPa

in II |ti I I I H I I I I I H n'11 in 11 m II M I II u i n
200 400 600 800 1000
-3
H o r i z o n t a l d i s p l a c e m e n t x 10" mm

89 Plot of load vs. horizontal displacement for samples from the toe of Moronga Park
' slump-earth flow (test 3, five repeated measurement, see appendix).

100

ft) 80
ft

CO
to
60-
CD
U
+J
CO
40
u
<0
CD
cn
20-

0 - ii i in i m i i n 11 n i in 11 II 11II11 II i m i l II "ii u ip i II in 111" in 11 II I "i n I N 11


10 20 30 40 50 60 70 80 90
Normal stress (KPa)

Fig. 8.10 Plot of shear stress vs. normal stress for Moronga Park slump-earth flow.
200-,

>
13
PH 81.3 3 KPa
CD 150
ft 67.7 KPa
54.1 KPa
oo 40.48 KPa
100
26.85 KPa

T3
C 50-
H
T3
cd
0
0 f 1111111111111111 M 111 n 11 1111111111111111111111111
0 200 400 600 800 1000
3
Horizontal displacement x 10" mm

(a)

160
81.33 KPa
- 67.7 KPa

54.1 KPa

~ 40.48 KPa

* 26.85 KPa

1
11111111111111 M 1111111111111111111111111111111111
0 200 400 600 800 1000
Horizontal displacement x 10"3 mm

(b)

Fig. 8.11 Plot of load vs. horizontal displacement for samples from Jetty rock slump, (a) test
1 from the toe of rock slump and (b) test 2 from the head of rock slump (five
repeated measurement for each test, see appendix).
200

150 81.33 KPa


CD
ft
67.7 KPa
00 *~ 54 . 1 KPa
100-

--. 40.48 KPa


>
H * 26.85 KPa
50-

0
0 fi i i ii i i ii i i ii ii ii i i ii i i ii i i i i i i ii i i i i i i i
0 400 800 1200 1600

Horizontal displacement x 10" 3


mm

Fig. 8.12 Plot of load vs. horizontal displacement for samples from the top of Jetty rock
slump (test 3, five repeated measurement, see appendix).

8 0 zi

~ 70
(0
ft
60:
CO
CO
CD
u 50: Test 1
CO
u
to 40:
CD

30:

20 II M nni[ II M i in ip u i m II |i m IIMI |i i in ini| l i m n u p II in in |i iiiiinri


10 20 30 40 50 60 70 80 90

Normal stress (KPa)

Fig. 8.13 Plot of shear stress vs. normal stress for Jetty rock slump.
200

81.33 KPa

*** 67*7 KPa

+ 54.1 KPa
*

' " ' ' 40.48 KPa

- 26.85 KPa
cd
0
0-fI I I II II I I| III I I II I I | I II I II II I| I I M II II I | I I I III I I II
0 200 400 600 800 1000
Horizontal displacement x 10"3 mm

(a)

too i
81.33 KPa
>
H * 67.7 KPa
(H

ft
54.1 KPa
rt
CO ~ 40.48 KPa
IN

T3
C
H

cd
0
iJ 0 fIIIIIIIII|IIIM IIII| IIIIIIII I|IIIIIIIIII IN IIIIII|IIIIIIIII |
0 200 400 600 800 1000 1200
3
Horizontal displacement x 10" mm

(b)
Fig. 8.14 Plot of load vs. horizontal displacement for samples from Harbou
test 1 from the top of rock slump and (b) test 2 from the toe of slump (five
repeated measurement for each test, see appendix).
81.33 KPa
.67.7 KPa
54.I KPa

40.48 KPa
26.85 KPa

i 1 1 1 1 1 1 1 i 1 1 ' ' '' I


r H n M"i M i " " ! Rnn T
600 800 1000
0 200 400 '""
Horizontal displacement x 1 0 " m m

Plot of load vs. horizontal displacement for samples from the head of Harbour
slump (test 3, five repeated measurement, see appendix).

80

cd
ft RO
bi

CO
CO $ = 16'
CD
u 40
4->
cn :
Test 2 C ^
u
to
CD
A 20
cn Test 3

0 IIIIIIII ill n n u n Mi n in in inn m II [ii in H n| m i n n ip n n u n pin in n |


10 20 30 40 50 60 70 80 90
N o r m a l S t r e s s (KPa)

8.16 Plot of shear stress vs. normal stress for Harbour slump.
<-v I6O-1
>

81.33 KPa
ft
OJ
120- 67.7 KPa
54.1 KPa

co
(N
40.48 KPa
> 26.85 KPa
H
T3
C
H

tj
cd
0
1
" 'I M I I 11 I I M I I II |l II II II II 11 I I I I n n 11 u u u | | 11 | | n n u !
) 200 400 600 800 1000 1200

Horizontal displacement x 10"3 mm

lOO-i

26.85 KPa
40.48 KPa
4J
C
CD 54.1 KPa
S
CD
CJ
to
-H -50-
ft 67.7 KPa
CO
H - ^ 3 ^ 8 1 . 3 3 KPa
a
CJ -100-
H
-U
U
CD
> 150'1111111111111111111111111111111111111111111111111111111111111
0 200 400 600 800 1000 1200

Horizontal displacement x 10"3 mm

Fig. 8.17 Plot of load and vertical displacement vs. horizontal displacement for samples from
the top of Coalcliff slump (test 1, five repeated measurement, see appendix).
250 q

>
H * 81.33 KPa
200-
cu
ft

150
co
CM

> 100
H

T3
cd
0 T T T T T ' I I I I I I I i rr\
W - ^ ~ ^ ^ 1000
- 3
Horizontal displacement x 1 ' ^

I6O-3
26.85 KPa
81.33 KPa
120:

*J
c
CD
&
CD
CJ
(0
iI
ft
CO
H
to
a
CJ
H
4J
CD
> - 8 0 - 11 1 1 1 1 1 1 1 i 1 1 1 1 1 1 1 1 'i 1 1 1 1 1 111 1 1 1 1 1 1 1 ' 11 1 1 1 ' ' ' i i' i 11 i
0 200 400 600 800 1000

Horizontal displacement x 10"3 mm

Fig. 8.18 Plot of load and vertical displacement vs. horizontal displacement for samples
from the head of Coalcliff slump (test 2, five repeated measurement, see
appendix).
160-1

>
H
81.33 KPa
120-
0)
ft 67.7 KPa
54.1 KPa
40.48 KPa
CO
80-
CN
26.85 KPa
>

40-
C

cd
0
0 tiniinu1iiiiiiiiUi'niiiii|iinniii|iiiiiiiii|MMiMii|
0 200 400 600 800 1000 1200
-3
Horizontal displacement x 10 mm

Fig. 8.19 Plot of load vs. horizontal displacement for samples from the toe of Coalcliff
slump (test 3, five repeated measurement, see appendix).

120-n

100
10
ft Test 2
ZL 805
CO
CO
CD 60
U Test 3
*J
CO
u 40
(0
CD
A
C/J
20:

0 1 1 1 ' 111 i < 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 J 1 1 1 1 1 ] 1 1 1 1 1 1 r 1 1 1 1 1 1 1 1 1 1 1 1 1 r~i J 1 1 1 1 1 1 1 1 1 1 1 1


10 20 30 40 50 60 70 80 90

Normal stress (KPa)

Fig. 8.20 Plot of shear stress vs. normal stress for Coalcliff slump.
160-1
>

u 54.1 KPa
CD
ft

CO

, t T 40.48 KPa
> ~- 26.85 KPa
H
Tl
C
H

cd
0
0 "ft II I I I I I I I " I H I I ! I M i l l I I I I I I ' ' " I
0 200 400 600 800 10UU

Horizontal displacement x 10"' mm

(a)

45 ->

40
to
ft
35
^"^
CO
CO
CD 30
u
<J
CO
u 25
(0
CD
A
cn
20:

15 |ni m i n|i11niin|nn111n 11 inn iii|i in M i n i m u m i| inn i n n


20 25 30 35 40 45 50 55

Normal stress (KPa)

(b)

Fig. 8.21 Plot of load vs. horizontal displacement (a), and plot of shear stress vs. norma
(b). Samples from highly weathered sandstone (niawarra Coal Measures) at the base
of Moronga Park slump-earth flow (three repeated measurement, see appendix).
cd
t
cd
a
it
cd
CX | e
cd *
CJ
X
00
<* f-H
c
uE-o o
o^
in o ^f 8 -a J 3Si ^xl <u
00 jt- ID CO c/3 O .
CO u C/3 TO c/a c3 -
N
C a S o u
CO cd co E > <- CO
g p fe B*>u
CO C u c c
u > .3
u o c>/3
CO
CO
l-t
a Q.
g I E-
& cd ^ CO c
P
O Q. c
tn u
H
(0
O -2 i-
X* cd -
S3
T3 P U 1)
ra ^- co
c
0
O C . P-S
w
E
c/> dj

H
LM
IJ
?.ico ^ co 3
e
J- J5C/3
0 M
ac - 2
nn Z .0T x ^usateoB T dsTp IEDTUSA. CL ^5
> CL
CN
CN
oo"
&b

o
-in (0

."1 C.
C
(U cn
g
<D
U w
(!)
l-H

a ro
g
m tj
(0 0
v z
c
0
N
11111111111111111 1111111111111111 |I I I I I I I I I | I I I I I I I I I | I I I I I I I I I | I II I I I I I I | I I I I II I I I | I I I I I 1 I I I | CN
H m o in o in o in
o o o o o 14
a oo co * CN *i- *t- n ro CN CN
0
(ATP asd N Z C ' T ) A f p ux p^oq ss (BdJl) S S 9 J 5 S JESI^S
60

50

40

X
0)
c
30

03
to
20

10
7

4
CL-ML
I l l l IIXl.i l I l l I II I I I I i ' i i i i i i I I I i i i i l I I I I I
10 40 50 60 80 90 10

Liquid limit wL

Fig. 8.23 Plasticity chart showing plasticity index (Ip) versus liquid limit ( W L ) for the matrices
of the Illawarra talus (between Clifton and Coalcliff). C L = inorganic clays of low
to m e d i u m plasticity. M L = inorganic silts and very fine sands with slight plasticity.

Sample localities are:


I. Selnei
2. Manglerud
3. Asrum
4. Labrador
S. Ottawa
6, 7. Sandnes
B. Little Belt
9. Bear paw
10. Pierre
II. Pepper
12. Cucharacha
13-IS. Vaiont
19. Walton W o o d
20, 21. Guildford
22-24. Acherfield
25, 26. Weald
27-28. Manglea
29. Wraysbury
30. London
31, 32. Gault
33. Chalk
34-36. Keuper marl
37. Liu
3840. Appalachian colluvium
39. Upper Coal Measures
04 OS 06
Residual strength, tt_

Fig. 8.24 Plasticity index plotted against residual strength coefficient (after Voight, 1973).
(T)Kenney(l9S9]
(7)'skempton-Gibson-BJerrum'ln Bjerrum and Simonl (I960)

(?) Hole (1942)

(7) Brooker and Ireland (I96S)

Mitchell (1965)

0Voljht(l973)

(7) Soil alone-peak values (<r = 0 l-OMkg/cm2)


ICanji
(?) Soil alone peak values (tr = 3 S leg/cm *) (1970,' 1972)

(V) Soil aloneminimum attained values (limited displacement)

MO) Soil-polished rock Interface

K 40
Plasclclty index. I

0 Bishop cfa(.(l97f)
.j A Townsend and Gilbert (1973)
fj. values* n Tulinov and Molokov (1971)
o Kan|l (1970, 1972), soil-polished rock

Fig. 8.25 Drained shear angle plotted against plasticity index. Figure shows the test results in
the form of curves obtained for peak strength under low and medium stress levels
(after Kanji, 1974).

2
C J n = 100-200 k N / m
40
- 48*
50
~
69
a.
-
>.43. 36- Happisburgh till
30 *
r'38* / London clay mixtures
1 1 i 1 1

4l^50 ^ / ' 05
CO

N ^ r 51
40 \Y
20
-- 0.4
20 N s
52 .^ s . Sand-benlonitz
0.3
\ N mixtures
33. 46 \ -3" /
9 . 11 12 /
"* \ v / 0.2
10" 1 5 ^ - JS 3^2.29 ^ " . . y
5
~ " 'V^?",A 32. -^
0.1
_
"I 1 1 1 1 1 I i i ! i i i i i i t . i 1 i i t t i t i M 1 i i t i i i i i i 1 i i i i r i t i i
20 40 SO B0 100
Plasticity index . I - (7.)

F i g . 8 . 2 6 Natural soils: residual friction angle against plasticity index (after Lupini et al., 1981).

CJn = 100-200 kN/m2


40 s

Sf so' H 0.6
XL
39. 0.5

40<
0.4
20'
0.3
i 7 .*
,,14,. 9 26 ."
3
S 10 . \6. 1 16 0.2
cn
ed * 5 " a\ 2'.27
or 0.1

1
" ' ' '" I ' " '" " ' I ' " 11 1111
20 40 60 80 100
Clay fraction ( 7. < 2 / u m ]

F i g . 8 . 2 7 Natural soil: residual friction angle against clay fraction (after Lupini et al., 1981).
26 ^

24

22:

J1 201
6
18
w
a. 16:

14-7

12 1 1 1 11 i 11 11 i 11 1 1 1 1 1 i i | 11 11 1 1 1 11 11 r i i 11 i 1 1 1 1 1 1 1 11 i I I |
24 28 32 36 40 44

Liquid limit ( W L ) %

Fig. 8.28 Relationship between the liquid limit and plastic limit for Illawarra talus.

22

20

+J 18-
c -
a.) ~
+J
c 16:-
o
o
+J
;
w 14:
o -
5 -"
12:

1 0 In II in HI 11 in in MI n in i n II in in i n n HI n HI in mil MI II in in II in i n i II
24 26 28 30 32 34 36 38 40

Liquid limit ( W L ) %

Fig. 8.29 Plot showing increasing moisture content with increasing liquid limit ( W L ) for
Illawarra talus.
2.20

2.00

.80 -
4-1
H
to
fl 1.60
CD

1.40 -
3

1.20 -

1.00 11 11 111 i I i M 11 II i i'| 'i n H I ITI ] 11 i "nil 11 11 11111 i i \'\~\ r n r n 1v\ vrrvvnTT"|
8.00 10.00 12.00 14.00 16.00 18.00 20.00 22.00

Moisture content (%)

Fig. 8.30 Relationship between bulk density and moisture content for Illawarra talus.
r
c. i

fl
o
H
4->
u
H
>-i
4I-t
1
fl
U
ID
4J
fl
H
4-1
0 .,- H
cn
rI
01
fl
to
11111111 . I .II , II i I i T T T i

10 15, "'h

Clay fraction (< 2 microns)%

Fig. 8.31 Decrease in angle of internal friction with increasing clay fraction for niawarra talus
(upper line for peak friction angle and lower line for residual friction angle).
50

G 40-

rH
01 30
fl
CO
fl
0 2.0-
rH
4J
U
H
l-l
4-1
10
tO
<D

T T T T T T T T I | I I I I I I I I I | 11 I I I I I I I | I I T T T I T T T T T T T T T T l T T r r r r i T T I I I | I I I I I F T T T ]
10 15 20 25 30 35 40 45

Silt and clay (%)


Fig. 8.32 Decrease in angle of internal friction with increasing percentage of clay and silt for
Illawarra talus.

30

CO
fl 25
o
l-l
u
H
e
CN
15
C
o
H
4-J
U 10
rO
u
cd

0-T
20 ""^''^ST'""^
Liquid limit (WL)%

Fig. 8.33 Relationship between clay fraction and liquid limit for Illawarra talus.
*

Stanwell Park

; ;;../ \ \ll I / Region C


'"". ..'/

Coalcliff.

C
cd
CD
0
o
Site 5
u
H
ti-l
< j Region B 0
cd
Jetty Pauit JJ

3
0
01

500 m

Legend

HawkesburySandstone
Narrabeen Group
Illawarra Coal Measures

Fig. 8.34 The study area site plan.


Fig 8.35 Plan of the Clifton earth slump.
sansai in UOHEASIH

o
O o o o
oo r- V3
m r- CN

CD
3

s
o
U

03

ttJ

u
rt
Cw
1 CJ
3 X!
S3
OJ *-
3
n C3
tlj

Fig 8.37. Plan of the Moronga Park slump-earth flow.
I Rnb 1 Bulgo Sandstone
IRnspj Stanwell Park Claystone
I Rns | Scarborough Sandstone
Wombarra Shale
Coal Cliff Sandstone
Bulli Coal
Pi IIllawarra Coal Measures

Fig. 8.38 The geology of the Clifton area showing the area affected by the Moronga Park
slump.

Clifton Fault

Moronga Park

Joint

Earth flow

Fig. 8.39 Geological cross-section of Moronga Park slump-earth flow above the cliff-
line shown in Fig. 8.37.
a
o
I
a
c
CD
H
> o
a
5
&
V
OJ

aa

00
CJ
2
o
a
o
Elevation in metres CO
I
</>
f/>

2 *
o
o cd

"O cU
CD
CO
r^
-S
ert e
Hi -*.
r-H ^
^v. o
CTj
Ui
C*t
O

i O
^t
OO
60
<-i
UH
(a)

(b)

F i g . 8.41 (a) PI*"1 f southern amphitheatre complex landslide (Graben A ) , (b) Aerial
photograph of the southern amphitheatre, this section of Lawrence Hargrave Drive (c)
has a history of the slippage, rockfall and mudslip.
(W-E) Water - filled joint

200
Rockfall /1J_^
j i
160
jj Perched water
L -120
Talus //-
SPC
Toppling
-80
Natural drainage
AA-A~\ I
Ft~T~l~\ I ~ 7 " SSs ,

Road ^ 7 / WSh 40
'Talusry_ Failure surface

Sea level *? :CSsr7T-rv:


irr^-^ Bulli Coal
0.00

Fig. 8.42 Generalised section through southern amphitheatre.


20 m Sea

Fig. 8.43 Plan of northern amphitheatre complex landslide


(Graben B ) .

Water -filledstress relief joint


(W-E)
ess . '
Natural drainage 90
CO

Rockfall CD

60 s
Natural drainage
Road
uau^ i
il'ii^H>? re
s-^yXj^.
a
o
WSh 30
Talus Slump //^T?^-~=^ - /
-Potential failure surface
a
Sea level ',.>: J i . .- ^CSs^ 0.00 >
f
,-. . :{ : ...^.v;. CD
U

Fig. 8.44 Generalised section through northern amphitheatre.


saijsai tn uon^A3'[3

o
o
o o
CN
o

CD CD
1-1
CO >
CO
CD Q
h
CO CD
73 >
Uj
CD
SO
-1 J3
1-1
BOJ
CD u
c3 <
=5 I<

-1

m
Fig. 8.47 Plan of the Harbour slump.
S3IJ9UI UI U0pTJA3J3

y>
71
CJ

CD

CO
CO
C
a
u
CJ

cd

o TJ
'""! CD
o cd
CD
CO
o
CO
CO
-a
,-J
O
) u
cd
rt
O CD
r>
DA) cd
O "^I

n 3
CJ CO
Si) CO
T3 -i
e; CTi
CO
,

cd
CD 1 )
-a
11
'
oo
*

oc
c3
(D
OX)
C/3 UJ

m
I6O-1
I (W-E)
J
150
140 -
130- 3 0 HJ
120-
110-
100-
Bulgo Sandstone \ g
90-

80- V BH 464
s BH 466
Road
70- *~\ .r-J-
Stanwell Park Claystone ^5^, ..... Tens i on t-.rack
e 60-
Water table j>^". Ta 1 us \^-____^
o 50- Indicated befe
a Failure surface Natural drainage
40- -
s 30-
20-
Scarborough Sandstone -A ~\-

10- Wombarra Shale

n- ^TusVQ^SealevelV

Fig. 8.49 Geological cross-section of Coalcliff slump (modified after State Rail Authority, 1982).
F i g . 8 . 5 0 (a) toppling in Scarborough Sandstone; (b) the toppling mechanism of failure is very
often induced by undercutting, Lawrence Hargrave Drive between Clifton and
Coalcliff.

Fig. 8.51 Significant rockfalls from the Scarborough Sandstone and Bulgo Sandstone bluffs
occur afterheavy rainfall between Clifton and Coalcliff along the Lawrence Hargrave
Drive. The patch is related to a big rockfall which necessitated repairs to the road by
Department of Main Road ( D M R ) .
Tension joints closer spaced wider open toward escarpment: edge

Bulgo
Sandstone
Water

Stanwell Park Claystone

(a)

Tension crack
' J...J-,i"
' Talus
!
Rockfall \ If Bulgo I Sandstone
(Slump or Topple) -Joints'! |
Stanwell Park
Water Claystone"

(b)

Fig. 8.52 (a) creep, setting and tilting of sandstone blocks on the claystone,(b) undercutting
(weathering and erosion) produces topple, rockfalls.

Toppling Fracturing

Fig. 8.53 Failure modes due to marine undercutting, between Clifton and Coalcliff, north of the
Coalcliff beach and north of the Stanwell Park beach.
Fig. 8.54 Significant work has been carried out by the Department of Main Road (DMR)along
the Lawrence Hargrave Drive including concrete retaining walls, shotcrete, gabions,
rockbolts and steel mesh to prevent falls from the side of the road. Lawrence Hargrave
Drive between Clifton and Coalcliff.

(b)

F i g . 8 . 5 5 (a) destruction of shotcret by swelling of the rock mass (Wombarra Shale); (b)
subhorizontal drains just below the contact between the Wombarra Shale and
Scarborough
TABLES TO CHAPTERS 1-3
Table 1.1 Typical D/L % Ratios for various landslide types based on data in

Skempton (1953), Selby (1967) and East (1978).

Landslide type D/L %

Flows 0.5-3.0

Slides 5-10

Slumps 15-30

Table 1.2 Causes of mass movement (after Terzagi, 1950).


External change in stability conditions:

1. Geometrical changes (undercutting, erosion, stream incision,


artificial excavation leading to change in slope height, length or
steepness).
2. Unloading (erosion, incision, artificial excavation).
3. loading (addition of material, increase in height, etc.) including
undrained loading (Hutchinson, 1970).
4. shocks and vibrations (artificial earthquakes, etc.). Associated
processes:
a) liquefaction;
b) remoulding;
c) fluidization;
d) air lubrication;
e) cohesionless grain flow;
5. drawdown (lowering of water in lake or reservoir); and
6. changes in water .regime (rainfall, increase in weight, pore
pressure).
Internal changes in stability conditions:

1. Progressive failure (following lateral expansion of fissuring and


erosion).
2. Weathering (freeze-thaw, desiccation, reduction of cohesion,
removal of cement).
3. Seepage erosion (solution, piping, etc.).
Table 2.1 Summary of averase whole rock component abundances in percent (after Odins et
al, 1990).

Stratigraphic Total Volcanic


Unix Lichics Pock Fragments Quartz Feldspar Matrix Cerent: Porosity

Loddon Sand-
stone Msaber 50.2 5.7 15.0 3.A 18.0 12.2 1.2
Lawrence Sand-
stone Msafaer 47.5 5.0 10.8 2.7 22.0 15.9 1.1
Nrvice Sand-
stone Member 41.2 6v0 10.1 4.7 24.0 20.0 -
Kenhla Sandstone 55.9 3.8 8.6 1.8 18.1 15.0 0.6
AllansCreek
Formation 43.9 5.8 12.6 4.4 8.3 30.8 -
Darkes Forest
Sandstone 47.8 12.3 9.2 1.5 15.4 25.9 0.2
Wilton Formation 22.0 0.5 37.7 1.5 16.4 19.9 2.5
Marrangaroo
Conglomerate 27.1 6.4 46.7 0.4 11.9 5.7 8.2
Erins Vale
Formation 43.0 23.2 20.8 1.1 18.5 15.4 1.2
Pheasants Ifest
Formation 61.5 16.8 2.7 6.6 10.5 18.1 0.6
Table 3.1. S u m m a r y of the general stratigraphy of the area from North Stanwell Park to Clifton
Fault (after Ghobadi and Pitsis, 1993).

GENERAL MAPPING THICKNESS


G E O L O G I C A L UNITS DESCRIPTION SYMBOLS (m)

Hawkesbury Massive Quartz


Mid-Late Sandstone Sandstone with minor Rh 100 +
Triassic shale beds

Newport
Formation Sandstone and
(Gosford interbedded shale Rhu 30
Formation)

Mid Bald Hill Red-brown claystone,


Triassic Claystone minor quartz lithic Rnbh 15
sandstone

Bulgo Quartz lithic sandstone,


Lower Sandstone minor shale and minor Rnb 120
Triassic C conglomerate
u
Stanwell Park Red-brown and greenish 37 - 53
c
0) Claystone claystone, minor quartz Rnsp
CD lithic sandstone
-C
C3
fc Scarborough Quartz lithic sandstone
Sandstone (coarse grained), minor Rns 26
shale & minor conglomerate

Wombarra Grey shale with quartz


Late Shale lithic sandstone (mainly Rnw 36
Permian fine grained)

Coalcliff Lithic quartz sandstone


Sandstone with minor shale Rnc 10 - 12

Bulli Coal Coal Bulli Coal 1- 3

Eckersley Dark grey carbonaceous shale Piy 20


Formation over light grey sandstone
C/3
<D
Balgownie
a;
1
cc Coal Seam
<D
C3
Sandstones overlying shales
C
coal and interbedded 34
u sandstone and shale

u.
Wongawilli Coal, mudrock and tuff Wongawilli
Coal bands, carbonaceous shale 9
C3 Coal
below

Based on Hanlon (1953) and Adamson (1974) and confirmed by currently available borehole data.
Thickness is somewhat variable. Those quoted are typical.
TABLES TO CHAPTER 4
Table 4.1 Point count analysis (with mineral proportions expressed as percentage) of the Coal
Cliff Sandstone, (a) fresh core samples from North Cliff, borehole IL55, (b) slightly weathered
samples and (c) moderately weathered samples between Coalcliff and Clifton. Note : K =
kaolinite, Cal = calcite, Sid = siderite and Ir.o = iron oxide.

Samples No. FCSs2 FCSsI FCSs4 FCSs3


Depth (m) 442.3 441.8 440 435.75
Quartz 30.5 29.4 26 31
Feldspars 8 3.6 10 4.5
Chert 5.5 9.4 4 9.5
Rock fragment 9 10 5 11
Calcite 17.5 3.2 23.3 1.5
Kaolinite 10 23.4 5 17
Iron oxide 8 8 4 2.7
Accessory. M 0 0.4 0 0.2
Mica 1 0.6 0 2
Siderite 0 0 12.3 1
Matrix 10.5 12 9.6 19.5
Porosity 0 0 0.2 0.1
Cement Cal + K k Cal + Sid K
(a)

Samples No. SWCSs5 SWCSs6 SWCSs7 SWCSs8 SWCSs9 SWCS10 SWCSsI


Quartz 26.2 24 26.5 22.8 23.2 24.5 25
Feldspars 3.2 5.2 2 2.4 1.3 0.5 1.6
Chert 8 4.5 10 6.8 7 9.2 5.6
Rock fragment 17.5 20 15.5 15.4 22 12.2 25.3
Calcite 27.5 1.7 1 24.4 5.3 0.5 0.9
Kaolinite 3.7 20.2 10.5 2.6 15.5 26.2 6
Iron oxide 0.7 6.5 14.5 8.2 7 14.2 12
Accessory. M 0 0 0 0.2 0 0 0
Mica 0 0.7 0 0 0.6 0.7 0
Siderite 0 0 5 0 4.3 0 0
Matrix 13.2 17 14.5 17.2 13.6 12 19.6
Porosity 0 0 0.5 0 0.2 0 4
Cement Calcite K + Ir.o K + Ir.o + Sid Calcite K + Ir.o K + Ir.o Ir.o + K

(b)

Samples No. MWCSS12 MWCSsI


Quartz 18 18.5
Feldspars 2.5 6.5
Chert 4 10
Rock fragment 19.7 14.2
Calcite 3 7.7
Kaolinite 17.7 10.2
Iron oxide 10.5 9.2
Accessory. M 0.2 0.2
Mica 0.2 1
Siderite 0 0
Matrix 22.2 22.5
Porosity 2 0.3
Cement K +Cal Cal + K

(c)
Table 4.2 Point count analysis (with mineral proportions expressed as percentage) of the
Scarborough Sandstone, (a) fresh core samples, West Cliff, borehole IL55, (b) slightly weathered
samples and (c) moderately weathered samples between Coalcliff and Clifton. Note : K =
kaolinite, Cal = calcite, Ir.o = iron oxide.

Samples No. FSSsI FSSs4 FSSs2 FSSs3


Depth (m) 394.7 396.8 400.5 402.1
Quartz 41.6 52.3 50 58
Feldspars 0.3 3 0.6 3.6
Chert 2.8 14.8 22.3 8.9
Rock fragment 14.3 8.3 11.6 6
Calcite 16 9 2.3 11.3
Kaolinite 11 1 2.3 0.6
Iron oxide 4 0.6 3.3 3
Accessory. M 0 0 0.3 0
Mica 0 0 0 0
Siderite 0 0 0 0
Matrix 9 8 4.3 6.6
Porosity 1 3 0 2
Cement Cal + K (Calcite Cal + K Cal + K + Ir.o
(a)

Samples No. SWSSs5 SWSRfifi


Quartz 32 32.3
Feldspars 1 1.6
Chert 1.3 2
Rock fragment 18.8 19.3
Calcite 5.6 3
Kaolinite 14 20.9
Iron oxide 6 9.3
Accessory. M 0 0
Mica 1 i1
Siderite 0
Matrix -18.3
o
10.6
Porosity 2 n
Cement Cal + K K+lr.o
(b)

Samples No. MWSSs? MWRsX


Quartz 26 30.6
Feldspars 1.5 3
Chert 15 2.6
Rock fragment 15 15 1 <~?
Calcite 3.3 3
Kaolinite 24.6 22.5
Iron oxide 8 9
Accessory. M 0 n
Mica o 0
Siderite o 0
Matrix 20.3 13.3
Porosity 0 1
Cement K+Cal K + Cal
(c)
Table 4.3 Point count analysis (with mineral proportions expressed as percentage) of the Bulgo
Sandstone, (a) fresh core samples, West Cliff, borehole IL55, (b) slightly weathered samples
between Coalcliff and Clifton and (c) moderately weathered samples at Stanwell Park Station.
Note : K = kaolinite, Cal = calcite.

Samples No. FBSs3 FBSsI FBSS2 FBSs4


Depth 290 298.4 306 316
Quartz 31 59 20 48.3
Feldspars 0.6 0.3 3.3 0.6
Chert 11.6 6 2 4.6
Rock fragment 7.6 10.3 17.3 10
Calcite 27 2.3 4.3 9.6
Kaolinite 12 10.6 29.3 12.6
Iron oxide 2 1.3 1.3 1.3
Accessory. M 0 0 1.3 0
Mica 2.2 0 0 1.4
Siderite 0 0 2.5 0.3
Matrix 6 10 20 9.3
Porosity 0 0.2 0 2
Cement Cal + K K K + Cal K+Cal

(a)

Samples No. SWBSs5 SWBSs6


Quartz 40.3 34.3
Feldspars 2.3 0.3
Chert 1.6 1.3
Rock fragment 15 13.3
Calcite 10.6 12
Kaolinite 12 5
Iron oxide 1.3 2.3
Accessory. M 0.3 0
Mica 0.6 2
Siderite 0 0
Matrix 16 28.6
Porosity 0 0.9
Cement K Cal+ clay
(b)
Samples No. MWBSs7 MWBSs8
Quartz 22.3 19
Feldspars 0.5 0.5
Chert 1 3
Rock fragment 10.6 13
Calcite 9 10.6
Kaolinite 31.3 18.3
Iron oxide 1.3 5
Accessory. M 0 0
Mica 1 3.6
Siderite 0 0
Matrix 23 27
Porosity 0 0
Cement k + Cal K + Cal

(c)
Table 4 4 Point count analysis (with mineral proportions expressed as percentage) of the Otford
L L o n e Member, (a) fresh core samples, West Cliff, borehole ^ \ % ^ % ^ ^
samples between Coalcliff and Clifton beside the road south of Jetty Fault. Note . Cal - calcite,
Sid = siderite.

Samples No. FOSM1 FOSM2 FOSM3 Samples No. SWOSM4 SWCONG5


Depth 414.7 412.5 411.5 Quartz 29 21
Quartz 27 26.3 26.6 Feldspars 1 2.3
Feldspars 2.1 2.3 2 Chert 11 9.3
Chert 15 15 6.3 Rock fragment 9 14
Rock fragment 17 17 9.6 Calcite 6 48
Calcite 21 21 20 Kaolinite 1 0.5
Kaolinite 2.3 1.3 1.9 Iron oxide 2 1.3
Iron oxide 1.3 2.3 6 Accessory. M 0 0
Accessory. M 1 0 0 Mica 0 0
Mica 1.1 2.3 2 Siderite 1 0
Siderite 0.2 0.5 2.3 Matrix 40 3.6
Matrix 12 12 23.3 Porosity 0 0
Porosity 0 0 0 Cement Clay Calcite
Cement Cal + Sid Cal + Sid Cal + Sid
(b)
(a)

Table 4.5 Classification of the Otford Sandstone Member, (a) fresh core samples, West Cliff,
borehole IL55, (b) slightly weathered samples between Coalcliff and Clifton beside the road south
of Jetty Fault.

Samples 1 2 3 Samples 1 2
Depth 414.7 412.5 411.5 Rock OSM OCONG
Q 27 26.3 26.6 Q 29 21
F 2.1 2.3 2 F 1 2.3
Chert 15 15 6.3 Chert 11 9.3
RF 17 17 9.6 RF 9 14
%Q 44.19 43.40 59.78 %Q 58 45.06
%F 3.44 3.80 4.49 %F 2 4.94
%RF 52.37 52.81 35.73 %RF 40 50
100 100 100 100 100
(a) (b)
Table 4.6 Classification of the Coal Cliff Sandstone, (a) fresh core samples, North Cliff,
borehole IL55, (b) slightly weathered samples and (c) moderately weathered samples between
Coalcliff and Clifton.

Samples 1 2 3 4
Depth 442 442 436 440
Q 29.4 30.5 31 26
F 3.6 8 4.5 10
Chert 9.4 5.5 9.5 4
RF 10 9 11 5
%Q 56.1 57.5 55.4 58
%F 6.87 15.1 8.04 22
%RF 37 27.4 36.6 20
100 100 100 100

(a)

Samples 1 2 3 4 5 6 7
Q 26.2 24 26.5 23 23.6 24.5 25
F 3.2 5.2 2 2.4 1.3 0.5 1.6
Chert 8 4.5 10 6.8 7 9.2 5.6
RF 17.5 20 15.5 15 22 12.2 25.3
%Q 47.7 44.7 49.1 48 43.8 52.8 43.5
%F 5.83 9.68 3.7 5.1 2.41 1.08 2.78
%RF 46.4 45.6 47.2 47 53.8 46.1 53.7
100 100 100 100 100 100 100

(b)

Samples 1 2
Q 18 18.5
F 2.5 6.5
Chert 4 10
RF 19.7 14.2
%Q 40.7 37.6
%F 5.66 13.2
%RF 53.6 49.2
100 100
(c)
Table 4.7 Classification of the Scarborough Sandstone, (a) fresh core samples W e s t Cliff,
borehole IL55, (b) slightly weathered samples and (c) moderately weathered samples between
Coalcliff and Clifton.

Samples 1 2 3 4
Depth 394.7 400.5 402.1 396.8
Q 41.6 50 58 52.3
F 0.3 0.6 3.6 3
Chert 2.8 22.3 8.9 14.8
RF 14.3 11.6 6 8.3
%Q 70.51 59.17 75.82 66.71
%F 0.51 0.71 4.71 3.83
%RF 28.98 40.12 19.48 29.46
100 100 100 100

(a)

Samples 1 2 Samples 1 2
Q 32 32.3 Q 26 30.6
F 1 1.6 F 1.3 3
Chert 1.3 2 Chert 1.5 2.6
RF 18.8 19.3 RF 15 15
%Q 60.26 58.51 %Q 59.36 59.77
%F 1.88 2.90 %F 2.97 5.86
%RF 37.85 38.59 %RF 37.67 34.38
100 100 100 100
(b) (c)

Table 4.8 Classification of the Bulgo Sandstone, (a) fresh core samples, North Cliff, borehole
IL55, (b) slightly weathered samples between Coalcliff and Clifton and (c) moderately weathered
samples at Stanwell Park Station.

Samples 1
Depth 298.4 305.5 289.6 316
Q 59 20 31 48.3
F 0.3 3.3 0.6 0.6
Chert 6 2 11.6 4.6
RF 10.3 17.3 7.6 10
%Q 78.04 46.95 61.02 76.06
%F 0.40 7.75 1.18 0.94
%RF 21.56 45.31 37.80 22.99
100 100 100 100
(a)
Samples Samples
Q 40.3 34.3
Q 22.3 19
F 2.3 0.3 F 0.5 0.5
Chert 1.6 1.3 Chert 1 3
RF 15 13.3
RF 10.6 13
%Q 68.07 69.72
%Q 64.83 53.52
%F 3.89 0.61
%F 1.45 1.41
%RF 28.04 29.67
%RF 33.72 45.07
100 100
100 100

(b) (c)
Table 4.9 X R D analyses of upper Illawarra Coal Measures
(highly weathered sandstone).
Location of samples : Clifton area (Moronga Park slump)

Samples Qz Feld Mc K
VSS1 A R F M
VSS2 A R F M
VSS3 A F F
VSS4 A C R
VSS5 A F T
R a w intensity
A= abundant > 500 Qz= quartz
C = c o m m o n 200-500 Feld= feldspars (orthoclase, plagioclase)
M = moderate 100-200 Mc ; = mica (muscoviteJ, sericite)
F= fair 50-10( C = carbonate (Ca, I% Ba)
R= rare 20-50 K = kaolinite (dickite, nacrite)
T= trace < 20 S = smectite
I:= illite
G:= goethite

Table 4.10 X R D analyses of Coal Cliff Sandstone (weathered shale


interbeds).
Location of samples : between Coalcliff and Clifton (road down
south of old adit)

Samples Qz Feld Mc C K G
HICSS1 M R T
LICSS2 M R
LICSS3 M R
HICSS4 M R
HICSS5 M R R
R a w intensity
A= abundant > 5 0 0 Qz= quartz
C = c o m m o n 200-500 Feld= feldspars (orthoclase, plagioclase)
M = moderate 100-200 M c = mica (muscovite, sericite)
F= fair 50-100 C = carbonate (Ca, Mg, Ba)
R= rare 20-50 K = kaolinite (dickite, nacrite)
T= trace <20 S = smectite
I = illite
G = goethite
Table 4.11 Wombarra Shale (moderately weathered samples)
Location of samples : Coalcliff area (beside Jetty Fault)

Samples Qz Feld Mc C K S I G
WS2 A F M T R
WS3 A R F T M R -
WS4 A F M R
WS5 A - R - M T R

R a w intensity
A= abundant > 500 Qz= quartz
C = common 200-500 Feldl= feldspars (orthoclase, plagioclase)
M = moderate 100-200 Mc = mica (muscovite, sericite)
F= fair 50-100 C == carbonate (Ca, Mg, Ba)
R= rare 20- 50 K == kaolinite (dickite, nacrite)
T= trace < 20 S == smectite
l := illite
G == goethite

Table 4.12 X R D analyses of Wombarra Shale (fresh samples).


Location of samples : West Cliff (borehole IL55)

Samples Qz Feld Mc K
WS1 A F R M T M
WS2 A F R M T
WS3 A F - M -
WS4 A F T M T
WS5 A M R M T

R a w intensity
A= abundant > 5 0 0 Qz= quartz
C = c o m m o n 200-500 Feld= feldspars (orthoclase, plagioclase)
M = moderate 100-200 M c = mica (muscovite, sericite)
F= fair 50-100 C = carbonate (Ca, Mg, Ba)
R= rare 20-50 K = kaolinite (dickite, nacrite)
T= trace <20 S = smectite
I = illite
G = goethite
Table 4.13 X R D analyses of Wombarra Shale (moderately weathered samples)
Location of samples : between Coalcliff and Clifton (north of
Jetty Fault).

Samples Qz Feld Mc C K S I G
WSHJ1 M R R
WSHJ2 M R T T
WSHJ3 C . R
WSHJ4 F F R
WSHJ5 M T R

Raw intensity
A= abundant > 500 Qz= quartz
C = common 200-500 Feld= feldspars (orthoclase, plagioclase)
M = moderate 100-200 Mc = mica (muscovite, sericite)
F= fair 50-100 C = carbonate (Ca, Mg, Ba)
R= rare 20- 50 K = kaolinite (dickite, nacrite)
T= trace < 20 S = smectite
I = illite
G = goethite

Table 4.14 X R D analyses of Wombarra Shale (moderately weathered samples)


Location of samples : between Wombarra and Coledale

Samples Qz Feld Mc K
WSHWC1 M . . . - F
M . . . R R
WSHWC2
M . . . R R
WSHWC3 F - - R R R
WSHWC4 R - - R R R
WSHWC5
R a w intensity
A= abundant >500 Qz= quartz
C = common 200-500 Feld= feldspars (orthoclase, plagioclase)
M = moderate 100-200 Mc = mica (muscovite, sericite)
F= fair 50-100 C = carbonate (Ca, Mg, Ba)
R= rare 20-50 K = kaolinite (dickite, nacrite)
T= trace <20 S = smectite
I = illite
G = goethite
Table 4.15 X R D analyses of Scarborough Sandstone (whole rock
highly weathered samples).
Location of project: Scarborough Station

CO \-
Samples Qz Feld Mc C K
SSS1 A - R M -
A T F M T -
SSS2
SSS3 A - F M F M
SSS4 A F F R M F F
SSS5 A R F F T R

Raw intensity
A= abundant > 5 0 0 Qz= quartz
C = common 200-500 Feld= feldspars (orthoclase, plagioclase)
M= moderate 100-200 Mc = mica (muscovite, sericite)
F= fair 50-100 C == carbonate (Ca, Mg, Ba)
R = rare 20-50 K == kaolinite (dickite, nacrite)
T= trace <20 S == smectite
I = illite
G = goethite

Table 4.16 X R D analyses of Scarborough Sandstone (weathered interbedded


grey shale).
Location of samples : Scarborough Station

Samples Qz Feld Mc K
SSSS1 A M F M F
SSSS2 A F - M F
SSSS3 A M - F R
SSSS4 A F - F F R
SSSS5 A F - M R

Raw intensity
A= abundant >500 Qz= quartz
C = common 200-500 Feld= feldspars (orthoclase, plagioclase)
M= moderate "100-200 MC:= mica (muscovite, sericite)
F= fair 50-10C C == carbonate (Ca, Mg, Ba)
R= rare 20-50 K = kaolinite (dickite, nacrite)
T= trace <20 S == smectite
l =illite
G = goethite
Table 4.17 X R D analyses of Stanwell Park Claystone (fresh
samples)
Location of samples : West Cliff (borehole IL55)

Samples Qz Feld Mc C K S I G
SPC1 A - M M R .
SPC2 A - F - M F .
SPC3 A - F - M F .
SPC4 A - F - F F .
SPC5 A - F - M F -

Raw intensity
A= abundant > 500 Qz= quartz
C = common 200-500 Feld= feldspars (orthoclase, plagioclase)
M = moderate 100-200 M c = mica (muscovite, sericite)
F= fair 50-100 C = carbonate (Ca, Mg, Ba)
R= rare 20- 50 K = kaolinite (dickite, nacrite)
T= trace < 20 S = smectite
I = illite
G = goethite

Table 4.18 X R D analyses of Stanwell Park Claystone (fresh


samples).
Location of samples : north of Coalcliff Station

Samples Qz Feld Mc K
SPC1 C F R M M F M
SPC2 M T R M M F -
SPC3 C T - F M F T
SPC4 C R F F M R -
SPC5 C
R R M F M
Raw intensity
A= abundant > 500 Qz= quartz
C = common 200-500 Feld= feldspars (orthoclase, plagioclase)
M = moderate 100-200 Mc = mica (muscovite, sericite)
F= fair 50-100 C = carbonate (Ca, Mg, Ba)
R= rare 20- 50 K = kaolinite (dickite, nacrite)
T= trace < 20 S = smectite
I = illite
G = goethite
Table 4.19 X R D analyses of Stanwell Park Claystone (moderately weathered
samples).
Location of samples : between Coalcliff and Clifton (Harbour
Fault)

Samples Qz Feld Mc K
SPCH1 F M H
SPCH2 M - R
SPCH3 M - R
SPCH4 F R R F
SPCH5 M R R

Raw intensity
A= abundant >500 Q z = quartz
C:= common 200-500 Feld= feldspars (orthoclase, plagioclase)
M== moderate "100-200 M c = mica (muscovite, sericite)
F= fair 50-100 C = carbonate (Ca, M g , Ba)
R=: rare 20-50 K = kaolinite (dickite, nacrite)
T= trace <20 S = smectite
I = illite
G = goethite

Table 4.20 X R D analyses of highly weathered Bulgo


Sandstone samples).
Location of samples : Stanwell Park Station

Samples Qz Fed Mc K
BSS1 A - R
BSS2 A F - F
BSS3 A - F R T M
BSS4 A - F -
BSS5 A - F -

Raw intensity
A= abundant > 5 0 0 Q z = quartz
C = common 200-500 Feld= feldspars (orthoclase, plagioclase)
M= moderate 100-200 M c = mica (muscovite, sericite)
F= fair 50-100 C = carbonate (Ca, M g , Ba)
R = rare 20-50 K = kaolinite (dickite, nacrite)
T= trace <20 S = smectite
I = illite
G = goethite
X R D analvsesof
Table 4 21 weathered grey shale interbedded in Bulgo Sandstone.

Location of samples : Stanwell Park Station

Samples Qz Feld Mc K
S1BSS A F F F
S2BSS A - C - F R
A R - - F T M
S3BSS
S4BSS A - M - M F
S5BSS A F M
Raw intensity
A= abundant >500 Qz= quartz
C = common 200-500 Feld= feldspars (orthoclase, plagioclase)
M= moderate 100-200 Mc = mica (muscovite, sericite)
F= fair 50-100 C = carbonate (Ca, Mg, Ba)
R= rare 20-50 K = kaolinite (dickite, nacrite)
T= trace <20 S = smectite
I = illite
G = goethite

Table 4.22 X R D analyses of fill


Location of samples: Coalcliff area (beside Harbour Fault)

Samples Qz Feld Mc K
H1 A - F M
H2 A - F F M
H3 A T R M
H4 A - F R M M
H5 A - F M
H6 A T R M M
H7 A - R M M
H8 A - F T M T
H9 A - R M M
H10
A - F M
M
R a w intensity
A = abundant >500 (3 z = quartz
C = common 200-500 Feld= feldspars (orthoclase, plagioclase)
M= moderate 100-200 Mc = mica (muscovite, sericite)
F= fair 50-100 C = carbonate (Ca, Mg, Ba)
R= rare 20-50 K = kaolinite (dickite, nacrite)
T= trace <20 S = smectite
l == illite
G = goethite
Table 4.23 X R D analyses of talus materials
Location of samples : Coalcliff area (Jetty Fault slump)

Samples Qz Feld Mc C K S I G
J1 A - C R F - - -
J2 A - C - M R - M
A - F R M - - -
J3
A - F R F T - R
J4
A T M - M R - -
J5
A - M T F R - -
J6
J8 A - - - M F F R
J9 A - C F M - C R

Table 4.24 X R D analyses of talus materials


Location of samples : Clifton area (Moronga Park slump)

Samples Qz Feld Mc K G
T1 A F T M T - R
T2 A R - - F R -
T4 A R - M F R
T5 A R - M - C R
T6 A R - M R -
T7 A F R M - R
T8 A R - M R -
T9 A F - M R -
T10 A R - M R . R

Table 4.25 X R D analyses of talus materials


Location of samples : Clifton area (beside Clifton Hotel)

Samples Qz Feld Mc K G
C1 C R R R F - - -
C1 A - R T M T M F
C3 A C R - M - - -
C4 A - F - C T M -
C5 A - R T M - - -
C6 A - R - M T - -
C7 A - F - M T . F
R a w intensity
A = abundant > 5 0 0 Qz= quartz
C = c o m m o n 200-500 Feld= feldspars (orthoclase, plagioclase)
M = moderate 100-200 M c = mica (muscovite, sericite)
F= fair 50-100 C = carbonate (Ca, Mg, Ba)
R= rare 20-50 K = kaolinite (dickite, nacrite)
T= trace <20 S = smectite
I = illite
G = goethite
TABLES TO CHAPTERS 5-6
Table 5.1 Main joint orientation foi lower Narrabieen urou p in the nc)rtnern i
area.

Formation Main Joint Orientation

CSs 005 035 065 135


CSs 015-0250 045 145
WSh 015 115 165
SSs 015 035 135 165
SPC 025 050 105
BSs 025 045 105
Bowman 005bc 055ac 105ac
c CM co rf in CD h- 00 CT O CM CO Tt in CD >- 00 CT O -- CO CT O T-
i CM CO cf in CD r
CO CO CO CO CO CO co CO CO "tf "cf "5f t "d" Tf "cf in in in in in m in in in in co co
CO
cc 5 "* <*

xz
c
o
E E
CO
E
^"
CD
ra
r^ CO CM -cf CT o CM CM T- 00 in sf CO CT in CO y- o "* r- CO CO CD r- *
T- in * C M
>
o "c lO cf "Cf CM CM o 00 h- h- N
*

CO CO CM h~ in CO co CM co in i- i-
in CT
"Cf CT
"Jf <a- "Cf sr 'Cf
m CT
"* sr Tf "Cf "Cf oCf CT
" CO r CO CO CO CO CM CM CM CM
E 'co m in LO in in
Ho co CO
3
E o
X "i_
CO CD
s D_ k_
o CO Tf ,_ CM CO -cr CO CO T CM o 1_ CO CM -cf CD 00 O CD CT o CM CO in CT CO "3- r-- -r-
CO CD CO CO CO CO CO in CO 00 CO CO r-- h- CO "Cf CD "Cf "Cf CO CO CO CO CO CD r- CO * in "cf
< *

CD OT CJ) CT CT CT CT CT CT CT CT CT CT CT CT CT CT CT CT OT CT CT CT CT CT CT CT CT CT CT CT

CM CO "cf in CO r- 00 CT O CM CO Tf m CD r-- 00 CT O CM CO "d" in CO r-- 00 CT O T-


CO CO CO CO CO CO CO CO CO "d"
5 -cr -5t "* rf "fr "d- "d" "d" in in in in in in m m in in co co
>> cc
xz o
c "E 10
o E, CO
00
E
E "co LO CO in X.
>*
CO ,_ CT in CO CO f- -a- CT CO CO "Cf CD in ,_ in co CO in in CM h- CO w n s
00 00 00 h- r- h- r^ CO CO CO CO m m Tf "t CO CM CM T T - CT CT CO 00 CM CM O c
cd E 'cto OCO CT)CO
CM OJ CM CM CM CM CM CM CM CM CM CM CM CM CM CM CM CM CM CM CM CM CM ,_ ,~ 1~ CO
o 'x TO
A
CO
o
in
2CO O
CD
CD
00 CO
< i_ o O CM in r-- -cf CO CD "* in CO CO CO I CO o "Cf T_ CO CT CM CO O CM co CM CT 1^ T - "d" x:
o CO h- CD CO in CD 00 r- -cf in 00 CO 00 CD 00 "tf "Cf CO i^- CO CO "Cf CD 00 CO CO CO h- in -t "cf D3
cd CD OT CT CT CT CT CT CT CT CT CT CT CT CT CT CT CT CT CT CT CT CT CT CT CT CT CT CT OTCTOT k-
3
> i
n
oo co
c c
0)
o CD
O
!-
I
D
Os
c
as c rriO-i-CMCO"cfincor^COCTOT-CMCO'cfincOh-COCTO'i-
cd ca T - w n t i o i D s c o
i
co u
'T-T-i-i-r-T-i-T-i-T-CMCMWCNWCMWC\ICMWnn
o xz
cc
CO c
as o
C E
CD "m
.2 CO
o g r ^ C M C T r r ^ ^ ^ ^ ^ ^ ^ ^ C O f ^ O T C M CO 1O- CO "d-CMTfh-COinOCDCOCT"S-0
ed !^^SCTT--i-CTCOh-COCOinCMi-Orf. CT . O O C_O C O T f C O C O C M -
i,- O_C T C O r -
*-> E '2 T T ^ r o o ) c > o o c o f f l o o c o c o c o o o c o s s s ( D i D ( 0 ( D ( o c D c o < o i o c D i n i n i n
cd E Q
C
O
c
Ic?
o co 3 S 2 2 f f l S r s f f i W w
r o i n T f s s s s m o n ^ ^ N m w c o i o S c Q f < J i n
c CD CTCTCTCTCTCTCTCTCTCTCTCTCTCTCTCTCTCTCTCT
o
cd
U >
cd
c
-a CO CNlCO"tlOCOSOOro^Wn^wcOls-COCOO'r-WPj'>flOCDSCOOjO'-
o '^,-1-1-1-T~-r-T--^"-CMCMCMCMCMCMCMCMCMCMCOC0
cc
U
cd if
cd
T3
o E
15 E ^ CO OC CM OiC-MCCTTCOTth-r C- OC iO tn +S TT tt S S + ^ - n n S S S ^ c s c n c o N C D c o i n u j n c M r
Ci-l
C
E r^cocoinm"d-^-"cf!$4 ^f"cf^^^f"cf"cfcOC0COCOCOCOCOCOCOC0C0C0C0C0
cd
1 I
CD
cU ra-
ce
-5 as CO CTCTCTCTCTCTCTC2CTCTc||c|c|gg^
.* as
TABLES TO CHAPTER 7
Table 7.1 Results of the durability tests for the lower Narrabeen Group (classification
according to Gamble, 1971).

No. Formation Weathering W% Id1% Id2 % Id3 % Id4 % Gr.N

1 BSs Fresh 1.42 98.61 97.83 97.19 96.53 HD

2 BSs SW 2.22 98.3 97.5 96.8 96.18 HD

3 BSs MW 1.72 94.6 90.83 87.58 84.34 MHD

4 SPC Fresh 3.2 93.48 81.08 66.43 54.77 MD

5 SPC Fresh 2.64 86.6 74.86 64 55.12 MD

6 SPC MW 4 61.14 16.31 5.46 1.16 VLD

7 SPC MW 6.32 27.52 9.94 4.85 2.81 VLD

8 SPC MW 2.03 30.17 3.45 1.01 0.3 VLD

9 SPC MW 3.6 50.87 25.12 21 18.6 VLD

10 SPC MW 3.28 91.27 70.6 54.5 41.3 MD

11 SPC HW 16.5 4.58 0.35 - - VLD

12 SSs Fresh 0.8 98.87 98.33 97.89 97.49 VHD

13 SSs SW 1.37 97.07 95.07 93.06 91.48 HD

14 SSs MW 2.76 93.4 88.6 85.42 82.25 MHD

15 OSM Fresh 2 98.52 97.64 96.89 96.29 HD

16 OSM SW 2.7 98.35 97.45 96.75 96.15 HD

17 WSh Fresh 1.51 98.6 96.-13 93.86 91.62 HD

18 WSh Fresh 1.43 98.3 96.89 95.56 94.36 HD

19 WSh MW 4.28 98 94 90.5 87.25 MHD

20 WSh MW 6 88 68 55.14 48.62 MD

21 WSh MW 6.16 87.56 68.5 55.68 48.4 MD

22 WSh MW 3.48 94 87.4 83.66 80.39 MHD

23 WSh MW 5.05 94.81 89.17 85.53 82.77 MHD


24 WSh MW 7.25 91 74.81 59.6 50.6 MD

97.4 96.85CSs
25 96.26 HD Fresh 1 98.2

95.22 94.73
26 CSs93.52 HD SW 1.97 97

27 CSs MW 1.25 96.24 93.6 91.55 89.73 MHD

S W = slightly weathered
M W = Moderately weathered
H W = Highly weathered
W % = Water content
Id1 % = Slake durability index (first cycle)
Id2 % = Slake durability index (second cycle)
Id3 % = Slake durability index (third cycle)
Id4 % = Slake durability index (fourth cycle)
Gr.N = Group n a m e
V H D = Very high durability
H D = High durability
M H D = Medium high durability
M D = Medium durability
L D = Low durability
V L D = Very low durability
Table 7.2 Result of the durability test for the lower Narrabeen Group (claystone interbeds in
the sandstone), classification according to Gamble, 1971).
No. Rock type Weathering W% Id1% Id2% Id3% Id4% Gr.N
1 Claystone (BSs) MW 10 0 0 - - VLD
2 Claystone (BSs) MW 14 1 0 - - VLD
3 Claystone (SSs) MW 7 91 74 61 55 MD
4 Claystone (SSs) MW 7 89 72 61 53 MD
5 Claystone (CSs) MW 3 94 86 77 71 MHD
6 Claystone (CSs) MW 2 97 93 90 87 MHD

BSs = Bulgo Sandstone


SSs = Scarborough Sandstone
C S s = Coal Cliff Sandstone
M W = Moderately weathered
W % = Water content
Id1 % = Slake durability index (first cycle)
Id2 % = Slake durability index (second cycle)
Id3 % = Slake durability index (third cycle)
Id4 % = Slake durability index (fourth cycle)
Gr.N = Group name
M H D = Medium high durability
M D = Medium durability
VLD = Very low durability

Table 7.3 Result of the durability test for the upper Illawarra Coal Measures (highly
weathered sandstone and Coal), classification according to Gamble, 1971).
No. Rock type Weathering W% Id1% ld2% Id3% Id4% Gr.N
1 Sandstone (ICM) HW 7 9 2 - - VLD
2 Sandstone (ICM) HW 8 13 3 - - VLD
3 Coal (ICM) Fresh 13 96 94 94 91 HD

ICM = Illawarra Coal Measures


H W = Highly weathered
W % = Water content
Id1 % = Slake durability index (first cycle)
Id2 % = Slake durability index (second cycle)
Id3 % = Slake durability index (third cycle)
Id4 % = Slake durability index (fourth cycle)
Gr.N = Group name
VLD = Very low durability
H D = High durability
Table 7.4 Slake durability classification (after Gamble, 1971).

% Retained after one % Retained after two


Group n a m e 10 - min cycle 10 - min cycle
(Dry Weight Basis) (Dry Weight Basis)
Very high durability > 99 >98
High durability 98-99 95-98
Medium high durability 95-98 85-95
Medium durability 85-95 60-85
Low durability 60-85 30-60
Very low durability <60 <30

Table 7.5 Slake durability classification (after Franklin and Chandra, 1971)

Slake durability (Id2 % ) Classification


0-25 Very low
25-50 Low
50-75 Medium
75-90 High
90-95 Very high
95-100 Extremely high

Table 7.6 Slake durability classification for Narrabeen Group strata ( classification according
to Franklin and Chandra, 1972).
The name of formation Weathering Id2% Classification
BSs Fresh 97.8 Extremely high
BSs SW 97.5 Extremely high
BSs MW 90.8 Extremely high
SPC Fresh 81.1 High
SPC MW 16.3 Very low
SPC HW 0.35 Very low
SSs Fresh 98.3 Extremely high
SSs MW 95.1 Extremely high
SSs HW 88.6 High
OSM Fresh 97.6 Extremely high
OSM SW 97.5 Extremely high
WSh Fresh 96.1 Extrem
WSh MW 94 Very high
CSs Fresh 97.4 Extremely high
CSs SW 95.2 Extremely high
CSs MH 93.6 Very high

S W = slightly weathered
M W = Moderately weathered
H W = Highly weathered
BSs = Bulgo Sandstone
S P C = Stanwell Park Claystone
SSs = Scarborough Sandstone
W S H = Wombarra Shale
O S M = Otford Sandstone Member
C S s = Coal Cliff Sandstone
>
T3
T3
o 3
TO
3 E
>j
E
B
IE
OT
a)
Q-l
E
5o
o
l_
CD
C\J co

o
-a
a
o
CJ
OT
C C
o
eU CD
a. a E
C/i
CD OJ|

2 >
J C
C o CO
CD
E
a CD
a CD
CD >.
o O
o i_

a.
00
Ui
OT
CO
jo CD
CU
c
t TJ
D. 3
C/3 CD
CO
>^

-a E CD"
OJ OJ
c C CO E
E
CC
C as
CO
o
CO
CD
XZo
5/5 >^o
CD >
D3| CD
cn
CD
ra
a CD
co
c
OS Q.
XZ cF g
H c x: 'co
CT
.2 ^o
B
o c
CD o c
D5| CD CO
TO
re TO
CO
E o'L. re
3 B o
CQ CD
o ui
"53
"O
SZ
OJ
I
sz sz
+' T
+-* Y^ T"' T~~ ^^ ^

b m m CQ b CQ
CD
CD CO
CO

JZ SZ
T T +-* T T~ T~ T~
Em -, ^ UI CO
CQ UI CD b m
LO
ro
CO
SZ .c y ^ y~ T~
T~
E 1
T T *b* QQ ~^~CO
Q
rn m Q CQ
ft
m
-cf

SZ X
T"~ T Y^ T~ T~"
E
^ -

OH ^^ CQ CO Q CO^~
CD CD Q CD b
CO
CO 3
OO
O
& c II CD x
CD ,_ T T -
T
<4-H
E
E CM < < Q CD O b CM
b m < D CQ
d, * - '

TJ c
H
TJ
3 CD SZ CU CD SZ
OT T
O OT E < CU a.
Q. < O m X
fc CQ < O CO
o UI
rn
^

^_
03 UI
,_ ,_ ,_
5 .,_ T
tl
c OJ < < UI O < cd
CM CQ < o<
U
CU
XuJ 5 ,_ ^ ^_
< UI < Q CD <
z 5
<
t
< UI CD <
T

l-l
a
o
s-i
CU TJ ,_ ,_ _ ,_ a Tl
^ ^. ,_ T_
> S-i CM < UI < Q CD <
CM < Q K o CQ < c/3
CU
.rH
Cd
CL,
cd
v. -a
-a i
JL. j T T ^
^^
cu x: 1 v T x~ Y y
ft < O < O CD < x ft < Q < o CD <
CM CM
00
CU
c/j t -
r->
y T~ T" T T""" X ~ 1
CU (
< O < a < < CD <
-t LO O CD < lO o o
a o
rH
o XZ
00
VH sz X
CD
< < CD CD < < < CD CO < CD

c/5
o CU
CJHITJ
s< o C
CD
C E
CD
S-l
.S3 r> o o CD
CU E TJ J E OT
"co CD c
ex oT < CD < CO < < o < CD < CD < < CO CD C
co '
II
T_ TJ o
CO CO o OT
r-i TJ O JX CO
cz U
"O W co sz co T; c
g O cn 2 co CO
XJ o E< CD < CD < <
cd * Q E< T-
CD < CQ < < E CL m
TV CO
2 Xco TJ 3=
CO =
g x rH
r3
VH

o o
CO g P n
fc O
3 S a) X X O 5= CO
CD m
TJ O
Q
> COo CO XZ ^ OT fa <+-? uJ OTO CO SZ ^ CO ?! o :5 O Oo
S5-3
a y
o
C/J n C/J UJ CO C/J
CD CO CO O o
3

a ^
C+H
C/J n C/J UJ CO C/J
CQ CO CO
O O
roII n H II
m CO OT
ft ^ CO
o o II II ro w CO CO
cc =2 oo
l-J -rJ
cd oo
rr OT O CO ^ oO
"
00 <U sz Crt M sz CO Q_
OO
s f Q
CU O)UJ
m lO r-~ CO in
a.
CD OT CM CO 00 r- CO
CO CO
OT CO
c
l> CO
a CD
r- "O i_ CO
CD 3
lO ft o ft CO o Ol o ft CO o
CU CD
CD T m o T co ft X oo CD CD m ft o ^ CO ft c O
3 ^ CD o
u E co ft ft ft ft Q. C) ft ^ ft ft 2T CD
- S cEc cd JU F E
E
cd H
CO
00 H cc
CO 5 E
H 1 SZ TJ
XZ sz
+^ y T-
*-^ ^ ^ ^- ^~
O b o O
b CJ
CO
CJ
CO o
X sz
t -r- T
^ ^ ^
EO O Eo o o
LO o LO
SZ XZ
1
^ T ^ ^ E CO
^ -
^ ^
E CQ CO UI
o CO UI
o
ft ft

X sz
T T
T T ^ ^ E CO ^ ^
CL, E CQ CO
a CO CL CO D CO
CO 3 CO
o O
a X
aa x
^ T
^
E ^ ^ ^ ^ E CO
^

CO Q CO < Q CO
a
cu CMm r o
flj CM
X X
71
*-M cd
SZ u- XZ
cd T y

Z E< T
<
^ ^ cd
O CO E <^ <
^ x
o CO
^
o 0
T
' z 0
1

01 o
'.
_. E <: T_ 1_ r_ -H E ,_ ,_ .,_
CU TJ 1
1) TJ 5 T
< < < < < <
cd CD CM
OT o cd 0 CM
OT o
riJ a
CO
CL CO TJ
UI CU UI
cd s ^- T ^ T 1 |_4 s T
! T T ^
r- T < < UI CQ < T < < UI CO <
C/3
i-H

=5 T3
13 ,
r *
TJ T
c/> TJ T -i
M T
fl cd CM < UI < D CO < CM < UI < Q CO <
CL,
r_, OH
cd
r- _ 00
O QJ CU ,
C/3
,_ flj i_

a)
c J_ T y
E J_ T T 1 ^
^ T r? T" T ^ ^

ft < Q < CQ < n ft < Q < CO <


CL
CM o rH
00 B
Tl CM O
00 oo OJ 00
TJ
TJ CL
00 TJ
TJ E i_ L_
C cd f- T " ^~ ^ ^~ T"" T"~* TJ E r ^ ^~ ^ T"" T"~ IT"
cd CO < m lO < < O CO <
C LO < O < o TJ o
>, o a B
wm C^H cd O
.v-*

XZ
cd uc X
< < CO CO <
T"

X
* i

U
<
T
< CO CD <
n
TJ tu O cd B
T
o
1)
UU tu
3
TJ
aj
r?
<u
3 X"' E T T
eu *-* E T
00 o < CQ < CO < < M CU oi < CQ < CO < <
C) V)
T cd X "
00oo
*1-*
fl) o <u
cd CL J
00 E' E< cd n, E<
-
T T T-
cd CO < CO < < Fi CO < CO < <
00 C/J cd
rj 00
TJ 0 TJ 0
r^ CU Q.
^
a.
OJ >H OT CJ CO X ^ CO flj'>< CO OT X ^ CO
ii X y CO a.CO UJ CO UJ ri>i X Y CO o0_ CO UJ CO UJ
X cd O CO CO CO <: O O CO CO CO O
cd CU O o X cd o
cd tu CC
o
H * rr 1 H r5
CL CL
3
XZ p E X
^i T T"" O "*~* ^
E 2 E
1
UI UI d
X
,1 jrf
E U UI
tU 00
tu X r^
^ _ ,_
cd H cd -? 1
CM Q Q cd OH CM Q Q
u, cd
cd&<
ZrE 5i r
L- flj
5?
. <u UI Q Q a *
tu
UI Q Q
o *
cd 00
<+H E
. cd TJ
Tl Tl
rH ._>
CM Q Q O TJ
0 CM Q Q
o
B oo OJ
J _H
0 E
* E3 b oo O
r- t_ rH TJ i
CL* Tl -1_ 7 T T- TJ
cd 0 X y-
1_ 7
cd E 0 ft Q O O ft Q O
- O OT CM
X
00 nCO CM o
to- "5 -^
*-> cd rn Ml
'^ OO UI > DO
<- X i_ _> 3 i
T~
r ^ ^~ ^ C/3 O f- y~ Y"~'
^~ ^~
B 2 to UI O o O lO UI O
.2r
flj L H
--> o o o
CO

tu n
rH
= -
L.
TJ E
T 2 3 i
T
Q O
CU o o $ r%
j j ^
Q O o o
TJ g S3 aj
S > E TJ g
CU -rH T
TJ
E T

Claystone
rH o O CO CO CQ o O CQ CO CO

andstone
rE <u C OJ
cd X
x E & OJ
cd o -5 2 0
a
UH *->
F T
F or
T T
!U *cj CO CO CO CQ cd co ^ CO
E >> CO CD CO CO CO

X
TJ
O "O
flj
TJ <>
O
TJ
C
CO
If
= o
anwe
arbor

OJ CO
0 X
00 M
n o
1o
a
r^ TJ
3 flj
0
Q.
CO
0
E
o
> OT co OT CO 5-. co co OT OT k_ OJ
UJ CO CO CO 3 .*: n n ?Jt CJ
iI T J
V o flJ
CO V CO CO CO CO E 3> 0 E CO w CO
O CQ CO o O CO CO C) o
flj
O '3 5J-g o h Xn CO
015 li II II II
c- <-< l
^ cu o II II II II f OT O co
CC 3 S3 cc UJ 0. CO
flj
r5rj -S E X TJ S b CO CO CO
o -g X cd
X cd
cd JJ
H
AS
0
o
>.
o CM CO
"55
OJ
CO o T d
o ^i
o^
X
OJ
0
%
X CM CM CD
c CO in in
o in CO in
in CM
X
E
CO CM CO
c m OJ
o in CD in
i^ in OJ
E
m
X
o> in
c
in CO CD
tu o r-- CO
CM
in
E
X
ft O) CD
c co CO CO
00 o CO CO
_flj
CL E lO CM
CO
X
E ft CO m ft
cd "E co CO
CO CD
o
oo o CM
00 in
OJ E
X
OJ co ft
CO
CD CM
o 1^ CO N-'
in T
E
E OJ
O XL co CO
ft
cd 0 r- N-' CO
DO 0 OJ
tu
5 T m
DO
DO CO cn
cd ft OT E
C/J v , > r^ ft O CD 0
TJ ft CO i^ o N-'
TJ in CM CM 0 o
TJ rJ CM O CJ
E 0 0
cd CU
co CD CO X -1
o ft **'

E ft OT
O d
,

X in CM
O
flj LO E O
ft CO lO
OJ 0
3
c CO CM
CM CM E
O X C
o 0
00 f~ o 3
' 1 0
CO ft ft O)
>* ;E 3 0
X ft O)
o. irf
CO CO
CO CD d OT E
0
TJ C/J
tu CU
CO
OJ CM CO o
00
T
E
^_
0
V
3
cd
o E
tu CO
ft
CM co
2
3
0
0 0
0
X
c/l E
O
C/3 H-H < oS in CO
c
X
3
E
O E
0
O
00 CM d 0 Q. CO
2 0
TJ CM T3 E
. E n) i_ 0 0 r o
X cd Q. 0 Q. 0 0 c (0
oo C/3 XL E CO TJ
m o
-*<
<U CL in LO F E
0 0 O X T)
ro CO To 0
3 sz O) o ft d co X W TJ 3 E
O
"_M
"5. in o
CO ft
ft
ft ft X
0 0 0 C
s-
0
CO 2
CO
3=
X CO o
*tf aE
Q
0 Hr-
o o o o
Mr-
X
r.
n
o
0 -t . OJ4^ 0 ^- rn
* fl) a X X X 3 O
[-- CU
X > CO CO CO OJ OJ
gj CO o Oo
OJ cd XL CO CO CO CO
0 CD 0
II
m
II II
5 5 5 -> n
X
cd cd
b
O CO CO
OO CO OT
OT
ll II ll CO CO UJ CO
HZ o < DO O CO CO O O
rr
Table 7.15 Point load strength classification for Narrabeen Group (fresh core samples, West
Cliff, borehole IL55), according to Franklin and Chandra, 1972).

Depth Formation Diametral ls(50) Axial ls(50) Anisotropy


From To Mean Range Strength Mean Range Strength la(50)
298.37 330.96 BSs 4.35 3.63-5.64 VH 7.25 5.96-8.41 VH 1.66
339.77 351.83 SPC 0.919 0.61-1.2 M-H 3.71 2.68-4.62 H-VH 4.04
396.79 402.11 SSs 4.37 3.60-5.76 VH 7.16 5.29-8.79 VH 1.51
411.9 414.7 OSM 2.78 1.24-3.68 H-VH 4.09 3.28-4.87 VH 1.47
427.3 436.6 WSh 1.69 0.71-2.89 M-H 4.4 3.27-5.19 VH 2.6
435.38 443.42 CSs 4.07 1.95-5.00 H-VH 5.53 4.98-6.57 VH 1.36

Table 7.16 Point load strength classification for Narrabeen Group (irregular samples, between
Clifton and Coalcliff), according to Franklin and Chandra, 1972).

Fomation Weathering ls(50) Strength


Range Mean
BSs MW 1.45-3.8 2.4 H-VH
BSs HW 0.41-0.9 0.65 M
SPC MW 0.23-0.92 0.5 L-M
SSs MW 1.42-4.56 2.38 H-VH
SSs HW 0.58-2.1 1.22 M-H
OSM MW 0.61-2.23 1.67 M-H
OCONG MW 3.55-5.93 4.28 VH
WSh MW 0.4-1.98 0.89 M-H
CSs MW 1.51-3.43 2.36 H-VH
CSs HW 0.41-1.64 0.99 M-H

BSs = Bulgo Sandstone


S P C = Stanwell Park Claystone
SSs = Scarborough Sandstone
W S h = Wombarra Shale
O S M = Otford Sandstone Member
O C O N G - Otford Conglomerate Member
M W = Moderately weathered
H W = Highly weathered

VH = Very high strength


H-VH = High to very high strength
M-H = Medium to high strength
M = Medium strength
L-M = Low to medium strength
Table 7.17 Summarised point load strength results for Coal Cliff Sandstone.

Degree of Orientation Number of ls(50) la(50) UCS


weathering of bedding samples MPa MPa MPa
Fresh Parallel 10 4.1 1.4 87
Fresh Perpandicular 10 5.5 134
MW 20 2.4 55
HW 20 1 23

Table 7.18 Summarised point load strength results for W o m b a r r a Shale.

Degree of Orientation Number of ls(50) la(50) UCS


weathering of bedding samples MPa MPa MPa
Fresh Parallel 10 1.7 2.6 40
Fresh Perpandicular 10 4.4 103
MW 20 0.9 21

Table 7.19 Summarised point load strength results for Otford Sandstone M e m b e r .

Degree of Orientation Number of ls(50) la(50) UCS


weathering of bedding samples MPa MPa MPa
Fresh Parallel 10 2.8 1.5 64
Fresh Perpendicular 10 4.1 96
SW(Congl) 20 4.3 99
MW 20 1.7 47

Table 7.20 Summarised point load strength results for Scarborough Sandstone.

Degree of Orientation Number of ls(50) la(50) UCS


weathering of bedding samples MPa MPa MPa
Fresh Parallel 10 4.7 1.5 108
Fresh Perpendicular 10 7.2 160
MW 20 2.4 56
HW 20 1.2 28

Table 7.21 Summarised point load strength results for Stanwell Park Claystone.

Degree of Orientation Number of ls(50) la(50) UCS


weathering of bedding samples MPa Mpa Mpa
Fresh Parallel 10 0.9 21
Fresh Perpendicular 10 3.7 ,83
MW 20 0.5 11

Table 7.22 Summarised point load strength results for Bulgo Sandstone.

Degree of Orientation Number of ls(50) la(50) UCS


weathering of bedding samples MPa MPa MPa
Fresh Parallel 10 4.4 1.7 104
Fresh Perpendicular 10 7.3 165
MW 20 2.4 56
HW 20 0.7 15
Table 7.23 Uniaxial compressive strength classification for Narrabeen Group rock (Based on
Geological Society Engineering Group Working Party Report, 1972).

Rock type Degree of weathering (UCS) M P a Term


CoalCliff Fresh Diametral 186.72 Strong
n
Sandstone Axial 133.52 Very Strong
II
Mean 110.12 Very Strong
MW 54.89 Strong
HW 22.9 Medium Strong

Wombarra Fresh Diametral 40.12 M. Strong


n
Shale Axial 102.73 Very Strong
II
Mean 71.42 Strong
MW 21.4 Medium Strong

Otford Fresh Diametral 64.18 Strong


n
Sandstone Axial 95.84 Strong
u
Member Mean 80 Strong
MW 46.9 Medium Strong
Otford Congl MW 98.73 Strong

Scarborough Fresh Diametral 108.11 Very Strong


n
Sandstone Axial 160.42 Very Strong
n
Mean 134.26 Very Strong
MW 56.42 Strong
HW 28.14 Medium Strong

Stanwell Park Fresh Diametral 21.14 Medium Strong


n
Claystone Axial 83.36 Strong
II
Mean 52.25 Strong
MW 11 Weak

Bulgo Fresh Diametral 108.11 Very Strong


n
Sandstone Axial 165.25 Very Strong
II
Mean 136.68 Very Strong
MW 56.42 Strong
HW 28.14 Medium Strong

S W = slightly weathered
M W = Moderately weathered
H W = Highly weathered
Table 7.24 Engineering geological study of fresh samples from the lower Narrabeen Group
in West Cliff borehole IL55.

No. of samples Depth (m) Formation %Quartz U C S (MPa) ls(50)


FCSsI 441.8 CSs 29.4 121.7 5.1
FCSs2 442.3 CSs 30.5 122.7 5.4
FCSs3 435.75 CSs 31 128.7 5.7
FCSs4 440 CSs 26 115.3 5.1
FOSM1 414.7 OSM 27 105.4 4.6
FOSM2 412.5 OSM 26.3 88.6 3.7
FOSM3 411.5 OSM 26.6 89 3.72
FSSsI 394.7 SSs 41.6 154.4 7.09
FSSs2 400.5 SSs 50 148.7 6.56
FSSs3 402.1 SSs 58 169.3 7.46
FSSs4 396.8 SSs 52.3 169.6 7.47
FBSsI 298.4 BSs 59 172.8 7.61
FBSs2 305.5 BSs 33 148.5 6.54
FBSs3 289.6 BSs 31 133.5 5.7
FBSs4 316 BSs 48 152.7 6.52

Table 7.25 Engineering geological study of slightly weathered sandstone samples from the
lower Narrabeen Group between Clifton and Stanwell Park.

No. of samples Rock type % Quartz U C S (MPa) ls(50)


SWCSs5 CSs 26.2 65.1 2.82
SWCSs6 CSs 24 52.9 2.25
SWCSs7 CSs 26.5 59.67 2.55
SWCSs8 CSs 22.8 49.1 2.07
SWCSs9 CSs 23.2 51.3 2.18
SWCSS10 CSs 24.5 61.4 2.67
SWCSs11 CSs 24 60.3 2.54
SWOSM4 OSM 29 44.2 1.93
SWSSsI SSs 32 77.9 3.4
SWSSs2 SSs 32.3 79.5 3.42
SWBSsI BSs 40.3 79.6 3.53
SWBSs2 BSs 34.3 67.2 2.94

Table 7.26 Engineering geological study of moderately weathered sandstone samples from the
lower Narrabeen Group between Clifton and Stanwell Park.

No. of samples Rock type % Quartz UCS (MPa) ls(50)


MWCSs12 CSs 18 24.4 1.05
MWCSs13 CSs 18.5 31.4 1.38
MWSSsI SSs 26 30.7 1.34
MWSSs2 SSs 30.6 34.1 1.47
MWBSsI BSs 22.3 24.9 1.09
MWBSs2 BSs 19 18.54 0.81
Table 7.27 Relationship between rock composition and durability for fresh and weathered
Wombarra Shale samples.

No. of samples Depth (m) Quartz Carbonate Clay mineral SDI


1. WS1F 437 304 21 78 96.1
2. WS4F 430 270 17 59 96.9
3. WSHJ1 220 56 87.4
4. WSHJ3 204 20 89.2
5. WSHJ4 161 84 74.8
6. WSHWC1 173 54 74.8
7. W S H W C 2 207 60 68.5
8. W S H W C 5 187 63 68

Note : this Table presents intensity of mineral data from XRD.


SDI = Slake durability index

Table 7.28 Relationship between rock composition and durability for fresh and weathered
Stanwell Park Claystone.

No. of samples Depth (m) Quartz Carbonate Clay mineral SDI


1.SPC1F 349 184 122 72 81.1
2. S P C 2 F 348 228 149 74.9
3. S P C 1 M W 135 169 386 9.94
4. S P C 2 M W 159 161 313 16.3
5. S P C 3 M W 81 52 136 74.9
6. S P C 4 M W 33 99 118 70.6

Note : this Table presents intensity of mineral data from XRD.


SDI = Slake durability index

Table 7.29 Relationship between rock composition and durability for weathered interbedded
claystone (CSs, SSs and BSs) samples.

No. of samples Quartz Carbonate Clay mineral SDI


1.LICSs1 196 28 85.91
2. HICSs5 234 55 32 92.64
3. S2BSs 277 102 0.2
4. S3BSs 153 159 0.15
5. SSSS1ST 278 83 74.31
6.SSSS2ST 271 75 72

Note : this Table presents intensity of mineral data from XRD.


SDI = Slake durability index
TABLES TO CHAPTER 8
Q CM CM CM ft r-- CM 00 O) CM in 00 OJ in m
UJ O) ft CM OJ CO CM ft T - LO ft o CO ft CO CM
CJ CM C\J CM CM CM CO CM CM CM CM CO CO 1_ T- CM

N r. 1 CM CO ft ft CM CO ft in CM 1^ CD in
cn CD ^ .O 00 ft O 00 LO o O) o
-5 C'j CO CM CO CO ft CO CM co CO CO ft CM i- CO

^ h- CO 00 CM in CO CO ft 00 CO m CD h-
CJJ CO m 1 OJ CM 1^ CO CM OJ CO 00 LO i
CO o o C'j r^ o t- 00 CD h- ft ft o
u+ CM CM CM CM CM CO T - T CM T CM ft T- "^ CM
CO
..
CO OJ CM CD CM LO 00 LO CO CO CM co 52
^ OJ O m
00 ft LO ,_ t- ft
u_ o
CJ) ft CO OJ CD m CO ft LO co CO *". CM
O)

6
CO
CO
0 _l _J _J _l l _i _i _l _i _l _i _J _l _l _l
o o O O O O O O o O O O o O o o
o
CO
CO CO ft ft CM ft ft CM CM CO ft ^ ^
_ J ft ft o UJ CO h- CO o 1^ CO LO CD in
-5 d ^

in m r-~ CO h- m CO LO LO 00 m CO Is* 'Ct


-e- CO CM r-- o co T CO t- 00 CD
, T
O) CO O 00 CM
'r" 'r~ CM CM CM ~ *" , ,_ T CM T - CM

QJ CM O) o ft LO CJ) 00 CM CM 00
71 y
m r- 00 ^ IV.
^9 T ft ft 00 ui 00 cj m m CO
CL, ~^~11 "*"T 1 "r~ "'"" 1 T

CU _J CO CO CO m 00 00 co CM ^ rv,
r- CL o LO ft CO CJ r-- CM O) 00 CM UJ co O -rl CM
, CM T CM T_ 1 l
CM CM \V. CM
o^ CM ~ ~ " CM
c/j
_l m in ft m m
TJ _ I CM ft T o 00 OJ ft m OJ
E CO CO CO C'j CO o
CO UJ CM CM CO CO CO
cd
--P co ft
ro
CM CM
0s-

a. in
in o CO 00 CO o 1^ in r-~ ft CO <tf CO CO
-e- ft ft CM CO 00 co ft C*J CO ft ft "r~ CO CM -i-
ft
u
LO LO
OJ
o ft co o o o CO 1"-. O CO
u o CM CM
CM b CM ft CM co CO CM
ft
cu
co
xCO ? r- r^ r- in CM ft CO co ft ft 00 h- T- CO c
.4 m in m T'.
CD CO i^ CO LO LO CO ft
3 r
^ 1
~
T~~
^ T 2
CJ
CrH

o
?-
CM CM CM r^
00 oo CO CO
m CM CO
m r- OJ
CM CM CO in CD ff
O) 00 OJ h- 00 o >
T
0
CO

CO CM OJ 0
flj TJ N
0 E
0 E E 'co
0
CU OJ 0 0
a, S. CO "fl- m i^. CO N. ft CD h>. LO [v. CD |v- 00 E E c 2 OJ
LO irt ft OJ t^ ft CO OJ |v- o 0 O i2
o* o T
E o o E
o E E ^ v-' E
DO 0 o oCM O
CO E .a 0 0 0 O
_E
I
0 O 0 E V 0 OJ N "H=
flj n E c CJ
k_
* iSDJ E .0
CO
E E ^ a. 0 0
U r.TJ
E 5 TJ 5 u 0 0 -5 TJ h_ 0 E E 0 0 TJ
5 5 u 0 . TT C.2
E
I
DO
0
CO 2
o
0
0
XZ
0
o 2o
0 0
0 o
SZ 2 0
CJ XZ
0
0 o 0
o ^o-^ XL
o 2 ^^ e ^
, s ^^,
o ^ CM roo ft LO
ll 0 >>^ E
CO -ti CD H - L_ r- 0
5 E CO C -S = t 3 h _
% | "i ~ = o %
CO
J E
0 oE
K 0
i: 0
O) TJ
5
E
6
- CM co ft in CD 1^ CO a> TJ 0 0 TJ
a 2 0 E o .
= ^ o 0 E 0 CO 0
0 II
0 C
c 0 1_
a=
3 J2 0 ~ 8 S.f-S Q-
= " - 0 E co
EII iSII
00 0 n TJ '- II CO
.o c
OJ
a, c XL a a S2
-1
o a Q. E n 0 CO
XL CJ n n
r3n 0 o 0 E 2 L- b O
E E 0 E Q- o. w
-,
CJ
~ O + CO
3 o O 3 0
3 OJ 3 0 3 O 3 o S- CO N
o a.
_ I O X (SiJL UJ ~3 O CO I CO
CL CO 6s- O CD
a
n o ff^ tf> J >
CO

H
Table 8.2 Plasticity index and estimated residual angles of internal friction for the talus in the
study area.
Location Sample Range Average Range Average
number Ip IP ^ <J>r $r
Clifton 1 15 21
Hotel 2 10 12 17.5 19
Slump 3 11 18.5
Moronga 4 14.5 21
Park 5 18.5 16 23.5 22
Slump 6 16 22
Jetty 7 11 18.5
Rock 8 10 10 17.5 17
Slump 9 9 16.5
Harbour 10 13 20
Slump 11 5.5 8 13 15.5
12 6 13.5
Coalcliff 13 11.5 19
Slump 14 14 14 20 17
15 17 22.5
tone - typical of
cd
3
TJ
"v\

typical of
CU
r,

esidual)
esidual)
u.
cd
aj

eak)
E
aj
rn rn co i DH ^^
o -r
13
cd c3 3
c2 3
<U cu U o 2 in
aj TJ
'co
E E
cn cn IH ^ 3 OJ
4 -.1 CL -2 rt r,

_2 -2 CO r , c3 < d< -H
co on E ^ O- cd
^, flj
OJ 4J ^ ^ a- ^ E ^, cd
CO
E
rn l_
TJ ^ E XZ cd
rE -^ rET.^ O C iS X ^
co c^ 3
o '55 TJ -
3 3
E O <+_ +_ 13 TJ
O O 5 2. <y >> o ^ O O
: E <U <U
3
TJ
'co
OJ
flj rH
E E cd 3 cd T J C \zz /-^ cd o o 'co
1*5 *- H-J H
.2.2 c aJ h u <u in r,
rE-E r,
o
" "H
E c ex > ft u EE cd
Ui 1 AJ
o o r-N . y \lZZy *^
E E
rr n OJ

13 13 2 p c t^ d
D u o
u u OJ O (U OJ
a. ex
C/J C/J aj <
c: s,U ETJ UE,CU in >^ y y a u o CJ
cd cd in X/,cd ,cd ,cd ,cd
O g r-
X -O CS,JcdE
> ,c2d ci: X
u
AJ i. 1- u, 3 o
Cr- C*r
l_ 1

cd cd Cd
"X 3 3
CJ "J CO 3
CO 3
CO
To <u.2
P - C <+- r
J ^T-
'E .E TJ CU u CU OJ
TJ b OJ \ZL > r- V- r-
5 E E E Ec2o
i
5 5 E _ 3 _ O ^1 cd
CU
c 3 a
<U flj

.3 'cd 'c3
3
'cd
op
'
3 _3
'c3 "c3
< C-H ., TJ TJ _ w C3JQ Cr- C+-.
CO cX cX X
XL r^
o
Ii I
> - . > >
JE
> .
<u J2
> .E
aj u2 >> > 1 >. 3 > >^ >-! o
>, >. >.
C/3 cd cd cd cd
o o o o o cd ~zz cd cd cd cd cd cd
CU
Q U.SU.SU UU U HUU U U UU
uu
r-

cd "13 ** 1 1 1
ON O N ^ ^ ^ i ^ ^ ^ i ^
NO ft NO ft 1 in 0 O r-H
r^ t-- oo NO ft
LO in m cn in in in cn cn
ci:

*&! 68 S& ^ ^ ^ c^ ^ ^ ^ ^ ^ ^
O 1 in
Eu oo o cn cn ~ m m C N in cn cn NO
cn cn cn CN ft cn in in cn CN cn

^^ ^_^ ^^ ^^
co y-s xs, s~-\
aj
**~\s~^s
^ 03 ^-> ^CQ CQ
CQ CQ CQ oo CU CQ CQ CQ
cQoo s00/
/-H" rr CO
s w w 00 w 00 1 1 "^ 00
sy
00
v
i i
" DO W
W
W ' *n ob -^m ^' in *^
CU ft^ "* ^ CN t> [>
T3, CN t^ ^ < cn r<
in r-* T-^
r< rH
CN 4i


^ CL,
"ex t >< CL
E r-H CN) cn ft m <' CN *t
-^ CN cn , 1 < ^-^. ^^.
r^ CN
X oP
rn
c5
00 cd U
aj *H
co in
CO flj

OJ r- r-

o o
CN CM
o
CN
o
CN
o
CN
1 r-1
CN CN
in
in
ON O N O N
CN CN CN
NO
CN
o o
E
cd
bo cn cn cn
aj
co
Q OJ
/v
XZ XZ
O
CQ
*H r-J
U rn CO
TJ 00 S3
E E E cd ^.^
Zo Zo ex 00 00 00 >-
Q
XL XI XL Xi XL <H >H_, Ho tt- U-. <H_
<*r <H_ <*H C*r
i*.
<H-H
E E E E E XI E H E ' ^ * *""' *^ ' *~' *^ E E 0H
OJ cd cd cd cd cd O JJ O o o o _o o O O
-O X X XJ XJ d <X X

cd d< cd cd cd cd cd
cd 13 H
cd cd cd
O O O
cd Id
*-^ *-^
TJ
E
^ oo
CS cj oou oo 00 00 o o o o flj
2 cu cU cu UU u UUU U U UU oxi
rJ
APPENDIX TO CHAPTER 7
Table 1. T h e result of Point Load Strength test for fresh Coal Cliff Sandstone.

Location of project: Coalcliff area

Specimens shape: Core

Depth : 436-443(m)

No Type W (mm) D (mm) P (MPa) De2(mnr) De (mm) Is F Is (50) U C S (MPa)


1 d 60 3 3600 60 1.11 1.08 1.19 27.19
M it
2 12 ll
4.45 4.8 109.02
II ll ll
3 4.9 1.81 1.95 44.34
it ll ll
4 12.5 4.63 5 113.43
n ll ll
5 13 4.82 5.2 118.09
ft ll ll
6 8 2.96 3.19 72.52
II II
7 13.5 5 ll
5.4 122.5
u
8 T.st
II
9 T.st
II ll
10 T.st
11 a 60 60 13.8 4583.66 67.7 4.02 1.14 4.58 103.9
u l
12 59 19.5 4507.2 67.1 5.77 1.14 6.57 148.53
n (
13 48 22.8 3666.92 60.5 8.3 1.08 8.96 204.07
u
14 50 15 3819.71 61.8 5.24 1.1 5.76 130.03
n
15 52.5 13.5 4010.7 63.33 4.49 1.11 4.98 112.62
n
16 60 15 4583.66 67.7 4.37 1.14 4.98 112.65
u
17 56 15.5 4278.08 65.4 4.83 1.12 5.4 122.89
u
18 T.st
u I
19 T.st
20 it 1
T.st

d = diametral Mean ls(50) // = 4.07


a = axial M e a n ls(50) 1 = 5.53
T.st = too strong la(50) = 1.4
-L = perpendicular
// =parallel
For abbreviations see chapter 7
Table 2. The result of Point Load Strength test for moderately weathered Coal Cliff Sandstone.

Location of project: Coalcliff area

Specimens shape: lump

2
No Type W (mm) D (mm) P (MPa) De (mnr) De (mm) Is F ls(50) U C S (MPa)
1 45 30 5.2 1718 41 4 0.91 3.65 84.7
2 35 30 3 1336 36.5 2.9 0.86 2.51 59.12
3 50 30 3.2 1909 43.7 2.2 0.94 2.07 47.62
4 50 32 5.5 2037 45 3.6 0.95 3.43 78.75
5 43 29 3.6 1587 39.8 3 0.9 2.71 62.89
6 40 35 3.8 1782 42.2 2.8 0.92 2.58 59.78
7 40 40 5 2037 45 3.2 0.95 3.05 70
8 40 40 3.5 2037 45 3.5 0.95 3.33 76.56
9 50 30 3.1 1909.8 43.7 2.16 0.94 2.03 46.75
10 50 34 3.6 2164 46.5 2.22 0.96 2.14 49.14
11 30 30 1.5 1145 33.8 1.4 0.83 1.45 34.65
12 28 30 2.5 1069.5 32.7 3.12 0.82 2.57 61.53
13 30 30 2.5 1145 33.8 2.9 0.83 2.43 57.75
14 32 40 3 1629 40 2.45 0.9 2.21 51.45
15 29 30 1.8 1107 33 2.17 0.82 1.79 42.91
16 35 37 2.2 1648 40.6 1.78 0.91 1.62 37.56
17 40 35 2.2 1782 42 1.64 0.92 1.51 35.01
18 40 44 3.5 2240 47 2.08 0.97 2.02 46.22
19 30 30 2 1145.9 33.8 2.33 0.91 2.13 46.4
20 34 40 3 1731 41.6 2.31 0.92 2.12 49.15

For abbreviations see chapter 7


Mean ls(50) = 2.36
Table 3. T h e result of Point Load Strength test for highly weathered Coal Cliff Sandstone.

Location of project: Coalcliff area

Specimens shape : lump

No Type W ( m m ) D e (mm) P (MPa) D e 2 (mm 2 ) De (mm) Is F ls(50) U C S (MPa)


u
1 52.5 32 3.3 2673.8 51.7 1.64 1 1.64 37.79
li
2 46 40 2 2342.7 48.4 1.14 0.98 1.12 25.61
ll
3 55 49 4.1 3431.38 58.57 1.59 1.07 1.7 38.55
n
4 50 22 1 1400.56 37.42 0.95 0.87 0.83 19.52
ll
5 48 35 1.9 2139 46.24 1.18 0.96 1.13 26.06
ll
6 54 37 2.2 2543.93 50.43 1.15 1 1.15 26.24
II
7 60 40 3.6 3055.77 55.27 1.57 1.04 1.64 37.16
II
8 50 30 1.9 1909.85 43.7 1.32 0.94 1.24 28.57
u
9 65 30 0.5 2482.8 49.82 0.26 1 0.26 5.9
n
10 52 32 2.9 2118.67 46.02 1.82 0.96 1.75 40.13
n
11 45 21 0.5 1203.2 34.68 0.55 0.84 0.46 11.03
M
12 42 39 1.2 2085.56 45.66 0.76 0.95 0.72 16.71
II
13 57 30 1.8 2177.23 46.66 1.1 0.96 1.06 24.38
II
14 60 31 1.1 2368.22 48.66 0.62 0.98 0.67 13.95
II
15 50 22 1 1400.56 37.42 0.95 0.87 0.83 19.52
II
16 55 35 2 2450.98 49.5 1.08 0.99 1.07 24.47
II
17 43 25 0.5 1368.73 37 0.48 0.87 0.41 9.828
II
18 53 28 1.2 1889.48 43.46 0.84 0.93 0.78 18.14
II
19 45 30 1.5 1718.87 41.45 1.16 0.91 1.06 24.65
II
20 40 25 0.5 1273.23 35.68 0.52 0.85 0.44 10.52

For abbreviations see chapter 7 M e a n ls(50) = 0.99


Table 4. The result of Point Load Strength test for fresh Wombarra Shale.

Location of project: Coalcliff area

Specimens shape : Core

Depth : 427-434 (m)

No Type W (mm) D (mm) P (MPa) D.2(mm2) De (mm) Is Is (50) U C S (MPa)


1 60 5.3 3600 60 1.96 1.08 2.12 48.02
ll
2 8.9 3.3 3.58 80.85
ll
3 5 1.85 2 45.32
ll
4 1.9 0.7 0.75 17.15
ll
5 1.5 0.55 0.59 13.47
II
6 4.5 1.66 1.8 40.67
ll
7 3.2 1.18 1.28 28.91
ll
8 5 1.85 2 45.32
ll
9 1.8 0.66 0.71 16.17
ll
10 7.2 2.67 2.89 65.41
11 60 63 10.8 4812.84 69.37 2.99 1.15 3.43 78.15
12 40 14.9 3055.77 55 6.5 1.04 6.76 153.56
13 50 13.5 3819.7 61.8 4.72 1.1 5.19 117.12
14 45 7.9 3437.7 58.63 3.06 1.07 3.27 74.23
15 45 10 3437.7 58.63 3.88 1.07 4.15 94.12
16 58 14.8 4430.87 66.56 4.46 1.13 5.03 114.39
17 65 10.5 4965.63 70.46 2.82 1.16 3.27 74.25
18 53 11.5 4048.9 63.63 3.79 1.11 4.2 95.26
19 50 10.9 3819.71 61.8 3.81 1.1 4.19 94.54
20 45 14 3437.74 58.63 5.43 1.07 5.81 131.73

d = diametral Mean ls(50) // = 1.69


a = axial Mean Is 50) J. = 4.40
J- = perpendicular la(50) = 2.60
// =parallel
For abbreviations see chapter 7
Table 5. The result of Point Load Strength test for moderately weathered Wombarra Shale.

Location of project: Coalcliff area

Specimen shape: lump

No Type W (mm) D(mm) P(MPa) D > r a ! ) De(mm) Is F Is (50) U C S (MPa)


1 72.5 35 2.5 3230.8 56.84 1.03 1.06 1.09 24.73
2 70 30 2.8 2673.8 51.7 1.4 1.01 1.41 30.88
3 70 35 2.9 3119.4 55.85 1.24 1.05 1.3 29.5
4 40 25 1.8 1273.2 35.68 1.88 0.86 1.61 38.05
5 80 26 2.2 2648.3 51.46 1.1 1.01 1.11 25.3
6 55 27 3 1890.8 43.48 2.11 0.94 1.98 45.59
7 70 31 2.2 2762.9 52.56 1.06 1.02 1.08 24.65
8 62.5 25 2.5 1989.4 44.6 1.67 0.95 1.58 36.41
9 65 30 1.2 2482.8 49.82 0.64 0.99 0.63 14.53
10 70 32 4.2 2852 53.4 1.96 1.03 2.01 45.75
11 25 35 0.5 114 33 0.59 0.82 0.49 11.66
12 25 40 0.9 1273 35.6 0.94 0.86 0.81 19
13 23 33 0.5 966 31 0.69 0.8 0.55 13.4
14 30 32 0.5 1222 34.9 0.54 0.85 0.46 10.85
15 16 30 0.5 611 24.72 1.09 0.72 0.79 19.97
16 28 40 0.5 1426 37.7 0.46 0.88 0.4 9.47
II ll ll il II n it it
17 it ll

II u ll ll it it n it
18 ll ll

II il ll II ll n ll it
19 ll li

ll n II N u n II it n it
20

For abbreviations see chapter 7 Mean Is (50) = 0.89J


Table 6. The result of Point Load Strength for fresh Otford Sandstone Member.

Location of project: Coalcliff area

Specimens shape : Core

Depth: 412-415 (m)

No Type W ( m m ) D (mm) P (MPa) De2(mnr) De (mm) Is F ls(50) U C S (MPa)


1 d 60 9.2 3600 60 3.41 1.08 3.68 83.54
ll u it ft
2 13.5 5.04 5.44 123.48
ll II ll ll
3 3.1 1.15 1.24 28.17
u u ll ll
4 10 3.7 3.99 90.65
n n li II
5 9.9 3.67 3.96 89.91
it II ll ll
6 11 4.08 4.4 99.96
it n ll ll
7 5.2 1.92 2.07 47.04
n n ll ll
8 4.1 1.52 1.64 37.24
it n
9 3.2 ll
1.18
ll
1.27 28.91
tl II
10 1.5 li
0.55
ll
0.59 13.47
11 a 60 50 15 3819 61.8 5.24 1.1 5.76 130.03
ll
12 50 11.9 3819 61.8 4.16 1.1 4.57 103.23
ll
13 52 8.9 3972.5 63 2.99 1.1 3.28 74.82
ll
14 56 11 4278 65.4 3.43 1.3 3.45 87.27
ll
15 56 11 4278 65.4 3.43 1.3 3.45 87.27
li
16 52 13.2 3972.5 63 4.43 1.1 4.87 110.86
It
17 50 7.8 3819 61.8 2.72 1.1 2.99 67.49
ll
18 51 12.5 3896 62.4 4.28 1.1 4.7 106.65
ll
19 49 11.5 3743.32 61.18 4.1 1.1 4.51 101.29
It
20 45 9.5 3437.74 58.63 3.69 1.07 3.94 89.52

d = diametral Mean ls(50)^ = 2.78


a = axial Mean ls(50)-L= 4,09
J- = perpendicular la(50) = 1.5
// =parallel
For abbreviations see chapter 7
Table 7. The result of Point Load Strength test for weathered Otford Sandstone Member.

Location of project: Coalcliff area

Specimens shape: lump

No Type W (mm) D (mm) P (MPa) De2(mm2) D e ( m m ) Is F ls(50) U C S (MPa)


1 60 35 3.2 2673.8 51.7 1.59 1.01 1.61 36.64
2 76 30 3.5 2902.98 53.87 1.61 1.01 1.66 37.71
3 55 40 4.2 2801.12 52.92 2 1.02 2.04 46.52
4 66 40 3.9 3361.35 57.97 1.54 1.06 1.63 37.18
5 65 36 4.8 2979.38 54.58 2.15 1.04 2.23 50.63
6 55 35 3.8 2450.98 49.5 2.07 0.99 2.04 46.91
7 60 37 4 2826.59 53.16 1.88 1.02 1.91 43.8
8 55 40 3.8 2801.12 52.92 1.81 1.02 1.84 42.1
9 73 40 12.2 3717.85 60.97 4.38 1.09 4.77 108.05
10 63 25 2.6 2005.35 44.78 1.73 0.95 1.64 37.77
11 40 35 1.6 1782.5 42 1.19 0.92 1.1 25.4
12 35 35 1.9 1559.7 39 1.62 0.89 1.44 33.73
13 35 25 1.8 1114 33 2.15 0.82 1.78 42.51
14 40 40 1 2037 45 0.65 0.95 0.61 14.21
15 35 45 2.6 2005 44.7 1.73 0.95 1.64 37.75
16 44 38 3.6 2128 46 2.25 0.96 2.16 49.61
17 37 30 1.8 1413.9 37.5 1.7 0.87 1.49 34.95
18 40 40 2 2037 45 1.31 0.95 1.24 28.65
19 40 34 8.5 1731 41.6 6.55 0.92 6.02 139.38
20 35 30 2.2 1336.9 36.5 2.19 0.92 2.01 44.64

For abbreviations see chapter 7 Mean ls(50) = 1.67


Table 8. The result of Point Load Strength test for moderately weathered Otford Conglomerate
Member.
Location of project: Coalcliff area

Specimens shape : Core

Depth : 412-415 (m)

2 2
No Type W (mm) D ( m m ) P (MPa) De (mm ) De (mm) Is F ls(50) U C S (MPa)
1 80 40 7.6 4074.36 63.8 2.49 1.1 2.74 62.66
2 68 36 7.9 3116.89 55.8 3.38 1.05 3.55 80.32
3 35 20 5 891.26 29.8 7.49 0.79 5.93 143.9
4 40 25 4 1273.2 35.68 4.19 0.85 3.59 84.82
5 56 38 7.5 2709.4 52 3.69 1.01 3.72 85.93
6 44 23 4.8 1288.5 35.89 4.97 0.86 4.27 100.79
7 67 38 11 3241.6 56.93 4.53 1.06 4.84.79 115
8 55 35 8.9 2450.9 49.5 4.84 0.92 3.61 109.5
9 46 32 5.9 1874.2 43.3 4.2 0.86 6.43 90.62
10 45 34 10 1946 44.1 6.85 0.94 4.24 148.7
11 70 32 8.8 2852 53.4 4.12 1.03 3.8 96.18
12 75 40 9.9 3819.71 61.8 3.46 1.1 4.8 85.85
13 69 32 9.9 2811.31 53 4.7 1.02 5.32 109.39
14 32 22 4.5 896.36 30 6.7 0.79 3.88 128.97
15 45 28 5.2 1604.28 40 4.32 0.9 3.73 90.72
16 60 42 8.5 3208.5 56.64 3.53 1.05 4.33 84.4
17 55 35 8 2450.9 49.5 4.35 0.99 3.74 98.5
18 72 35 8.6 3208.5 56.6 3.57 1.05 3.56 85.34
19 40 25 4 1273.23 35.68 4.19 0.85 3.85 84.82
20 68 40 7.8 2720 52.15 3.82 1.01 3.94 88.34

For abbreviations see chapter 7 Mean ls(50) = 4.28


Table 9 . The result of Point Load Strength test for fresh Scarborough Sandstone.

Location of project: Coalcliff area

Specimens shape : Core

Depth : 397-402 (m)

No Type W (mm) D (mm) P (MPa) De2(mnr) De (mm) Is F Is (50) UCS(MPa)


1 d 60 8.9 3600 60 3.3 1.08 3.56 80.85
i l n n
2 14.9 5.52 5.96 135.24
i 1 n ll
3 16 5.93 6.41 145.28
l l il ll
4 13.5 5 5.4 122.5
l i li ll
5 12 4.45 4.8 109.02
1 l ll ll
6 9 3.33 3.6 81.58
i l n ll
7 12 5.34 5.76 130.83
l i n II
8 12 4.45 4.8 109.02
l i n ll
9 8.4 3.11 3.36 76.19
it tt
10 i ti
10 3.7 4 90.65
11 <a 60 52 13 3972.5 63.02 4.37 1.1 4.8 109.37
12 46 18.2 2760 52.53 8.8 1.02 8.98 204.09
13 50 18.5 3000 54.77 8.2 1.04 8.52 193.39
14 47 18.2 2820 53.1 8.6 1.02 8.79 186.34
15 48 16.8 2880 53.66 7.79 1.03 8.02 182.21
16 49 13 2940 54.22 5.9 1.03 6.07 138.58
17 50 15.9 3000 54.77 7.07 1.04 7.36 166.74
18 47 13 2820 53.1 6.15 1.02 6.27 143.24
19 47 14.5 2820 53.1 6.86 1.02 6.99 159.78
20 42 10 2520 50.19 5.29 1 5.29 120.52

d = diametral Mean ls(50)^ = 4.37


a = axial Mean ls(50)i-= 7.16
_L = perpendicular la (50) = 1.5
/^parallel
For abbreviations see chapter 7
Table 10. The result of Point Load Strength test for moderately weathered Scarborough Sandstone.

Location of project: Coalcliff area

Specimen shape : lump

No Type W (mm) D(mm) P(MPa) De2(mra2) De (mm) Is F ls(50) U C S (MPa)


1 60 30 8 2291.8 47.87 4.66 0.984.56 104.04
2 50 33 7 1650 40.62 5.66 0.91 5.15 119.46
3 60 40 6 3055 55.27 2.62 1.04 2.74 62.01
4 60 40 7 3055 55.27 3.06 1.04 3.2 72.43
5 70 50 5 4456.3 66.75 1.49 1.13 1.69 38.26
6 60 40 4.5 3055.7 55.27 1.96 1.042.05 46.38
7 60 60 7 4583.6 67.7 2.03 1.142.32 52.46
8 60 50 10 3819.7 61.8 3.49 1.1 3.84 86.59
9 45 45 7 2678.3 50.7 3.49 1.01 3.49 79.82
10 65 35 6 2896.6 53.8 2.76 1.03 2.86 64.61
11 36 35 3 1604.2 40 2.49 0.9 2.25 51.82
12 33 43 2.5 1806.7 42.5 1.84 0.92 1.7 39.6
13 40 40 2.4 2037 45 1.57 0.95 1.49 34.33
14 40 35 3.2 1782.5 42.2 2.39 0.92 2.21 51.1
15 50 40 3 2546.4 50.4 1.57 1 1.57 35.82
16 40 35 2 1782.5 42.2 1.49 0.92 1.38 31.86
17 60 30 3 2291.8 47.87 1.74 0.98 1.7 38.93
18 50 35 3.4 2228.1 47.2 2.03 0.97 1.97 45.18
19 45 45 3.5 2578.3 50.7 1.81 1 1.81 41.39
20 45 40 2.5 2291.8 47.8 1.45 0.97 1.42 32.42

For abbreviations see chapter 7 M e a n ls(50) = 2.38


Table 11. The result of Point Load Strength test for highly weathered Scarborough Sandstone.

Location of project: Coalcliff area

Specimen shape: lump

No Type W(mm) D ( m m ) 2 2
P (MPa) De (mm ) De (mm) Is F ls(50) U C S (MPa)
u
1 58 35 2.9 2584.6 50.8 1.49 1 1.49 34.1
2 ,, 60 45 2.8 3437.7 58.63 1.08 1.07 1.16 26.2
n
3 60 36 2.8 2673.8 51.7 1.39 1.01 1.4 32.03
M
4 45 30 3 1718.8 41.45 2.33 0.91 2.12 49.52
H
5 52.5 38 2 2540 50.3 2.1 1 2.1 47.88
II
6 40 32 1.6 1629.7 40.37 1.31 0.9 1.17 27.6
II
7 55 35 1.7 2450.9 49.5 0.92 0.99 0.91 20.84
II
8 75 35 1 3342.25 57.8 0.39 1.06 0.41 9.4
II
9 55 35 1.1 2450.9 49.5 0.59 0.99 0.58 13.37
II
10 75 38 1.9 3628.7 60.2 0.69 1.08 0.74 16.92
II
11 60 35 2.4 2673.8 51.7 1.19 1.01 1.2 27.42
II
12 65 30 1.5 2482.8 49.82 0.8 0.99 0.87 18.17
If
13 55 35 1.2 2450.9 49.5 0.65 0.99 0.64 14.73
II
14 67 35 1.9 3241.66 56.93 0.78 1.06 0.82 18.69
15 " 62 32 3 2526.1 50.26 1.58 1 1.58 36
II
16 60 48 2.9 3666.9 60.5 1.05 1.08 1.13 25.81
II
17 43 32 2.8 1751.9 41.8 2.13 0.92 1.95 45.4
II
18 53 35 1.5 1855 43 1.07 0.93 0.99 23.03
II
19 48 30 3.1 1833 42.8 2.25 0.93 2.09 48.35
20 " 63 35 2.5 2807.4 52.9 1.18 1.02 1.2 27.44

For abbreviations see chapter 7 Meanls(50) =1.22


Table 12. The result of Point Load Strength test for fresh Stanwell Park Claystone.

Location of project: Coalcliff area

Specimens shape: Core

Depth : 340-352 (m)


2 2
No Type W(mm) D(mm) P (MPa) De (mm ) De (mm) Is F Is (50) U C S (MPa)
1 d 60 1.8 3600 60 0.66 1.08 0.712 16.17
2 n
60 2.5 ll
0.92 1.08 1 22.54
ll
3 II
2 ll
0.74 0.8 18.13
n II
4 II
2.5 0.92 1 22.54
n n
5 it
1.2 0.44 0.48 10.78
II II
6 n
3.8 1.4 1.52 34.3
n II it
7 3.1 1.15 1.24 28.17
it
8 n
3 il
1.11 1.2 27.19
n n n
9 2 0.74 0.8 18.13
n
10 it
1.5 li
0.55 0.6 13.47
11 a 60 40 4 3055.77 55.27 1.74 1.04 1.8 41.18
II
12 54 10.8 3240 56.92 4.45 1.06 4.71 106.62
13 II
40 7.1 3055.77 55.27 3.1 3.1 3.22 73.38
n
14 52 7 3120 55.85 2.99 1.05 3.13 74.41
II
15 50 5.8 3000 54.77 2.58 1.04 2.68 60.84
n
16 48 10.5 2880 53.66 4.86 1.03 5 113.67
n
17 47 9.5 2820 53.1 4.49 1.03 4.62 104.58
n
18 50 8.8 3000 54.77 3.91 1.04 4.06 92.21
II
19 39 6.8 2340 48.37 3.88 0.98 3.8 87.16
n
20 46 7.1 2760 52.53 3.43 1.02 3.43 79.55

d = diametral Meanls(50)<^= 0.91S


a = axial Mean Is (50) 1 = 3.71
J-=F >erpendicular la(50) = 4
^ =parallel
For abbreviations see chapter 7
Table 13. The result of Point Load Strength test for moderately Stanwell Park Claystone.

Location of project: Coalcliff area

Specimens shape: lump

No Type W ( m m ) D (mm) P (MPa) D;(mm 2 ) De (mm) Is F Is (50) U C S (MPa)


1 d 40 40 0 2037.18 45.13 0 0 0 0
i
2 42 35 1.4 1871.66 43.26 0.99 0.93 0.92 21.35
1
3 37 23 0.5 1083.52 32.91 0.61 0.82 0.5 12.05
t
4 42 36 1.3 1925.13 43.87 0.9 0.94 0.84 19.5
t
5 32 22 0.7 896.36 29.93 1.04 0.79 0.82 20
i
6 50 40 0.3 2546.47 50.46 0.15 1 0.15 3.42
t
7 38 25 0.6 1209.57 34.77 0.66 0.84 0.55 13.25
i
8 40 37 0.5 1884.39 43.4 0.35 0.93 0.32 7.55
1
9 46 30 0.2 1757 41.91 0.15 0.92 0.13 3.2
i
10 35 26 0.5 1158.64 34.03 0.57 0.84 0.47 11.37
11 ia 37 37 1 1743.06 41.75 0.76 0.92 0.69 16.19
i
12 38 35 0.5 1693.4 41.15 0.39 0.91 0.35 8.26
13 ' 40 28 0.5 1426 37.76 0.46 0.88 0.4 9.47
i
14 35 20 0.2 891.26 29.85 0.3 0.79 0.23 5.76
i
15 33 21 0.5 882.35 29.7 0.75 0.79 0.59 14.39
i
16 40 25 0.3 1273.23 35.68 0.31 0.85 0.26 6.27
t
17 42 25 0.2 1336.9 36.56 0.2 0.86 0.17 4
i
18 32 29 0.5 1181.56 34.37 0.56 0.84 0.47 11.2
i
19 36 25 0.5 1145.91 33.85 0.58 0.83 0.48 11.55
II
20 40 21 0 1069.52 32.7 0 0 0 0

For abbreviations see chapter 7 Mean Is (50) = 0.50


Table 14 . The result of Point Load Strength test for fresh Bulgo Sandstone.

Location of project: Coalcliff area

Specimens shape : Core

Depth : 298-331 (m)

No Type W ( m m ) D (mm) P (MPa) De2(mm2) De (mm) Is F ls(50) U C S (MPa)


1 d 60 9.1 3600 60 3.37 1.1 3.63 82.56
n
2 12.1 4.48 4.83 109.76
ll
3 10 3.7 4 90.65
ll
4 17.6 6.52 7.05 159.74
ll
5 12.5 4.63 5 113.43
It
6 9.8 3.63 3.92 88.93
ll
7 8.3 3.07 3.32 75.21
ll
8 14.1 5.23 5.64 128.13
ll
9 9.5 3.52 3.8 93.1
it
10 10 3.7 4 98
11 60 50 13.9 3000 54.77 6.18 1 6.42 145.75
12 50 15 3000 54.77 6.67 1 6.93 157.31
13 57 24.8 3420 58.48 9.68 1.1 10.35 234.58
14 58 19.8 3460 58.99 7.64 1.1 8.17 185.82
15 49 18 2940 54.22 8.17 1 8.41 191.9
16 55 17.8 3300 57.44 7.2 1.1 7.63 173.17
17 52 15.7 3120 55.85 6.72 1.1 7.05 159.75
18 50 16.1 3000 54.77 7.16 1 7.44 168.86
19 53 10 3180 56.39 4.19 1.1 4.4 100
20 50 12.9 3000 54.77 5.74 1 5.96 135.37

d = diametral Mean Is (50) // = 4.35


a = axial Mean Is (50) -L = 7.25
-L = perpendicular la (50)= 1.7
'^parallel
For abbreviations see chapter 7
Table 15. The result of Point Load Strength test for moderately weathered Bulgo Sandstone.

Location of project: Coalcliff area

Specimens shape : lump

No Type W (mrr ) D (mm) P (MPa) D e 2 (mm 2 ) De (mm) Is F Is (50) U C S (MPa)


tl
1 75 35 9 3342.25 57.8 3.59 1.06 3.8 86.56
it
2 72.5 45 7 4153.94 64.45 2.25 1.12 2.52 56.87
li
3 55 40 6.2 2801.12 52.92 2.95 1.02 3 68.61
n
4 67.5 40 7 3437.74 58.63 2.71 1.07 2.9 65.74
II
5 70 35 5.5 3119.4 55.85 2.35 1.05 2.46 55.86
n
6 80 35 5 3565 59.7 1.87 1.08 2.02 45.71
n
7 60 40 7.5 3055.77 55.27 3.27 1.04 3.4 77.4
u
8 55 45 10.2 3151.26 56.13 4.32 1.05 4.53 102.9
n
9 80 35 7.2 3565 59.7 2.69 1.08 2.9 65.75
it
10 60 35 5.9 2673 51.7 2.94 1.01 2.96 67.75
n
11 35 50 3.6 2228 47 2.15 0.97 2.09 47.78
it
12 35 40 2.5 1782.5 42 1.87 0.92 1.72 39.92
u
13 40 40 2.2 2037 45 1.44 0.95 1.37 31.49
n
14 40 45 3.2 2291.8 47.8 1.86 0.97 1.82 41.59
n
15 35 48 3.1 2139 46 1.93 0.96 1.85 42.55
n
16 35 40 2.8 1782.5 42 2.09 0.92 1.93 43.92
II
17 40 40 3 2037 45 1.96 0.95 1.86 42.86
18 35 37 3.3 1648.8 40.6 2.67 0.91 2.43 56.34
u
19 35 48 3.5 2139 46.24 2.18 0.96 2.09 48.15
II
20 35 50 2.8 2228 47 1.67 0.87 1.45 37.1

For abbreviations see chapter 7 Mean ls(50) = 2.4


Table 16. The result of Point Load Strength test for highly weathered Bulgo Sandstone.

Location of project: Coalcliff area

Specimens shape : lump

2 2
No Type W (mm) D (mm) P (MPa) De (mm) De (mm ) Is F Is (50) U C S (MPa)
1 40 25 0.1 1273.23 35.68 0.1 0.85 0.08 2.02
2 40 30 0.1 1527.88 39 0.08 0.89 0.07 1.66
3 60 40 1.6 3055.77 55.27 0.7 1,04 0.72 16.57
4 60 30 1.2 2291.83 47.87 0.7 0.98 0.68 15.6
5 60 40 1.9 3055.77 55.27 0.83 1.04 0.86 19.64
6 50 40 1.8 2546.47 50.46 0.94 1 0.94 21.46
7 30 30 0.1 1145.91 33.85 0.1 0.83 0.08 2
8 58 36 1.9 2658.52 51.56 0.95 1.01 0.95 21.87
9 65 40 2.3 3310.42 57.53 0.92 1.06 0.97 22.14
10 40 30 0.5 1527.88 39 0.43 0.89 0.38 8.95
11 50 32 0.1 2037.18 45.1 0.06 0.95 0.05 1.3
12 60 50 1.4 3819.71 61.8 0.49 1.1 0.53 12.15
13 60 35 1.8 2673.8 51.7 0.89 1.01 0.9 20.51
14 70 40 2.6 3565.07 59.7 0.97 1.08 1.05 23.71
15 60 40 1.2 3055.7 55.27 0.52 1.04 0.54 12.3
16 70 45 3.4 4010.7 63.33 1.13 1.1 1.24 28.34
17 40 30 1.2 1527.88 39.08 1.04 0.89 0.93 21.67
18 50 38 1 2419.15 49.18 0.55 0.99 0.54 12.43
19 65 30 2.5 2482.8 49.82 1.34 0.99 1.32 30.44
20 60 50 1 3819.71 61.8 0.34 1.1 0.37 8.43

For abbreviations see chapter 7 M e a n ls(50) = 0.65


APPENDIX TO CHAPTER 8
E X A M P L E O F DIRECT S H E A R TEST DATA
(Accordong to Bowles, 1992).

Location : Harbour slump


Sample No.: 1 Description of sample : talus
Sample dimensions :
Side: 6 cm
Ht: 3 c m
Area: 36 c m 2
Volume: 108.0 c m 3
Weight of talus used : 189 gr Bulk density ( i w ) = 1.75 g/cm3
W % = 13.5 Dry density ( y^ ) = 1.5 g/cm3
Normal load = 9.85 kg Normal stress = 26.85 KPa
Loading rate = 1.2 mm/min Loading ring constant = 1.87 N/div

Vert, dial Vert. Horiz.dial Horiz. Correction Load dial Horiz. Shear
reading displace. reading displace. area reading shear stress
(0.01mm) AV mm (0.001mm) AH m m A' N/div force (N) (KPa)
0 0 0 0 108.0 0 0 0
-2 -0.02 50 0.05 [ 12.5 23.3 6.4
-1.5 -0.015 100 0.1 ^ x 17 31.8 8.83
n
-1.5 150 0.15 ] 20 37.4 10.3
II
-1.5 200 0.2 / 23 43.01 11.94
if
-1.5 250 0.25 / 26 48.62 13.5
-3 -0.03 300 0.3 / 30 56.1 15.5
-8 -0.08 350 0.35 / 31 57.97 16.1
-10 -0.1 400 0.4 32.5 60.77 16.8
-6
-4
-0.06
-0.04
500
550
0.5
0.55
IJ
/
34
34
63.58
63.58
17.6
n

ll n
1 0.01 600 0.6 / 34
ll n
2 650 0.65 34
4
0.02
0.04 700 0.7
I
108.0 34 II ll

Note : plot on Figs using dial readings.


Load dial reading x loading constant = horiz. shear force
For example : 12.5 x 1.87 = 23.3 (N)
Horiz. Shear force (10) / area = shear stress (KPa)
For example : 23.3 (10) /36 c m = 6.4 ....etc.
For this sample in test 1 normal stress is 26.85 (KPa) and shear stress is 17.6 (KPa).
See figures 8.13 and 8.15 (test 1).
A copy of full data set is available in the Department of Geology, University of Wollongong.
APPENDDC - SAMPLE LOCATIONS
eoi-cMcoTtincoi^coo)Oi- CMco'd-incoi^oocjj O T - W C O ^ W C O S 00 CJ) O i - CM
i - T - CM CM CM
000)0)0)0)0)00)0)0)0)00 o o o o o o o o cocococococococo CO C O C O C O C O
wvninmininmwinmwcoco cocococococococo C M C M C M C M C M C M C M C M CM C M CM CM CM
CMCMCMCMCMCMCMCMCMCMCMCMCVI
COCOCOCOCOCOCOCOCOCOCOCOCO CMCMCMCMCMCMCMCM CO CO CO CO CO CO CO COCO CO CO CO CO
COCOCOCOCOCOCOCO
l-l-l-l-HI-l-l-HI-l-HI- l-l-HI-l-h-HH H H I- H H I- h- I 1 I I I

C O C O C O C O C O C O C O C O C O C O C O C O C O cocococococococo CO CO CO CO CO CO CO CO CO CO CO CO
x* x* i x" x" x" x" x" x" x" x"
x" x"
x" x" x I x"
x" x" j J I j I I
x" x" x" x" x" x" x"
x" x" x x" x" x" x"
CO
E
CD
a.
2
to

CD
CD C
o CD
co 4' c
c Tl CO
o CD
c TJ
to c
ca CO
m c TJ
CO o
E c to
n
LL CO
CO .
uCO SZ :
CJ) CO
TJ
c
o O o CO
o o O) CO
TJ
X> CQ
i_

CO o
CJ) o
CO
O
CM
CO
I
(-

CO
co

>s O o o O o o
o
o CO CO = = cfl = = CO CO
o o o
co o o O O
o <fl O O O
O CD

CO
CD CD CD CD CD CD
D c c C C CO CD
0) 0) CD CD CD CDCD c c CD CD
O o
CO CD
< c c c cz c c 1= C c C C o o c c
o o o o o o O CO CO o n TJ CO CO
CO CO CO TJ o o
CO
T3
CO CO CO CO CO (0
T3 TJ TJ -D T3 TJ TJ TJ c c 2 2
CD CD CD CD CD CD 1_ 1- m CD CD CD CD CD
c
co c co
co c co
c co
c ccoCcO CO CO CO CO CO CO to o)
c c c CO CO CO CO CO COCO TJ T J c c c CCOO CCOO TJ TJ 5 O O O CCO CO c c TJ TJ
CC CO CO CO
o CO "D o o o TJ TJ CD CD " CO CO " o
o
TJ
o
TJ TJ TJ CO
TJ T3 T3 T3 TJ TJ CD CD
CD CO 0) CD CD CD CO CD i i_ *-.
k
CD CD CD CD
* - . *-.
TJ
TJ T J ^* CD
2 2 c c
c C7 r. c i i i i i CO CO CO
CO i _ XL XL C C to CO CO CO

1
CO Cfl CO CO CD CD CD CI) CO CD CD
TJ TJ CO CO
+-. *>
TJ T J T J SZ SZ CO CO CD CO CO
CO CO CO CO r *-*
r i: r r r -C CO CO
' t * - ^ *-< -
C C C tS tS CD CD J= XL CO
CO CO TJ
XL CD CD TJ TJ
co co co co co coCO CO CO c c
CO CO CO
CD CD <D CO CD CO CD 55CO CO CO
CO CO -> ->
Ui Ui Ui CD CD S
CO CO
CO
CD 55 CO CO
CO
i
Q]
i_

X
CD
i
CD CD
* l_ 5 55 55 5 Z >< >
>. CD CO
LL LXL XL X
> CD CD
LL L L L L L L CO CO
CD CD
SZ sz
LU
LL L L L L >%>,>,>,>,>, CO CO CO

55 >,> $ >. >i
sz sz szCfl +. *I

-C sz j= .c -C .c sz * _ i SZ SZ
^^B B Cfl
CD CD CD
X
D) Ol Ol D) O Ol CJ) CD CD CD CO >< ^ B B
O) O) CO CD CO CO CO CD CD
^ f= TJ c fl CD
TJ CD CD
Ui Ui (Ji Ui Ui (IiCO O O x: xrTJ T J co co
o g) g) o
O O U_
o
1 1

LL. Li- 55
CO CO >> >
z O i- CM CO in co r* oo SZ
O) O)
o ft $ .. . w
CO co co co
cB cB co co rr\ m m rr> SET * *
CO CO
Ui Ui UiCO CO
CQ CO CQ CQBE?
X
K K KCO COKCO CO
I CO g;
O Og O O CO CO LL p p p
II
555 5 CO CO
3 incoNOfflOi-wntiocoN I - T - C M C M C M C M C M C M !ONcoo)Or-c\in *OOOJO-^CMCO^W m co s oo
O O O O O T 1 T-T 1 T - T - 1 c o c o c o c o c o c o c o c o C M C M C M C M C O C O C O C O co co co co co
a cococococococococococococo m m m m m m m i n
c o c o c o c o c o c o c o c o co
i n i o i n i a n i n i o m in
co
w
co co co
in m m
Q- l ox
x i oxu ixi oxi x
n ix
o ixo ixa i
xo x o mxm X X X X X X X X x x x x x x x x X
m ix X X X X
<
CO
g
UJ

8
CD
o

CD 0)
c c CO
CD o o
CO 10 CO
C TJ TJ
oO cO cO co
C C C
TJ CO CO
c
Cfl = sz = = x:
CO O) CD
3 "J
O ok_
O
k-
O)
o
XJ o
k n
CO CO cfl
o o
CO C/J

CD

CO CO
CO x:
I

X
21-
co o D)
3
CD
3
O =
CO-
D) to k- O
c CD
o i_
< o o
XJ
X)
1

2 cfl
o
<
CO CD 0) CD CD
X CO CD 0) CD CD CD CD CD CD CD CD C C c c
X C C c cfl CO CO C CO
C C
IT c r O O CO O O O O
10 COfl c
O
n o CO CO co O CO CO
CO CO CO
4- 4-1
4W 4
TJ TJ TJ CO CO
TJ TJ TJ TJ TJ 10 CO

I ^ n
TJ CD CD CD CD TJ TJ
r
CO cc Cr.
O C C C C C TJ TJ
fl CO CO L_ cnu. k- 0)
a) 0) _CD aj a)
CL CZ CZ CL C C
l_ CO CO CO CO CD CD Cl) O O O O Cfl CO CO CO CO CO CO c c
X TJ TJ TJ TJ r .c: r CO CO CO CO CO 4- 4-* * 4-< TJ CO
CO TJ CO
TJ CO
TJ CO
LU Q CI) CD OJ c cfl CO CO sz x: sz sz sz CO CO COCO CD CD CD CD co co 2co .ca
3?
CD CD ID a) CD
CD CD CD CO Ui Ui ui cn TJ TJ TJ
TJ L_ k_ CD CD
ca ca CO
ca Cfl
ca
CJ Ui c
XL cSZ
X 10
ID
SZ x: .r.
CO cfl CO CO
5>. >5* 5
>> sz sz sz X sz C C C C
CO CO Cfl CO
CD CD
XL X Cfl CO Ui Ui
c
X cX
CJ CD CD CD CD LL
CO CO
LL (0 LL
LL CO LL
CO CO CO CO CO CO ffl CD CD Ui Ui
p CD CI) CD CD CD CD CD TJ TJ TJ TJ
55 555 ca CO CO
CD
x: x: x: x:
CD CD

55 55
CD CD CD CD
>> >. k- \ k_ k
p X X sz sz
CO CD
TJ TJ
co co to co CO CD CD CD
X u> gj g> g O O
CO CD CD CD x: x: x: x: X X
k_ k. k_ k- >* >> at ra 4-< 4-<
4-* 4-<
LL X I X X LL LL LL LL
Q 2 2 x: x: X X co ca CO
CD
CO
CD
X g> o) CD CD

X X X 5 55 5
T- CM co * in CM co * in
5
o m cn Ui m m cn Ui cn cn CO CO co cS co
u> cn Ui cn Ui 5 5 5 5 5 5 5 5 5
> > > > >
I
LU
_]
0.
O) O T - CM CO
co * *
CO CO CO CO CO
m co

m in in in in m in m w
CO O ) O T - CM co * m
co co co co <* * m m w in m m
* *

co co co co co co co co
co
in
co
c~. OO
in m
co co
O) o
in
co
co
co
i-
CD
co
n ^ in
CO CO CO CO
co co co co co
ID s
CO

in m in m m m m m in m in
2 X X X X X X X X X
X X X X X
in in X X X Xm X in m m in m
X X X X X X X X X X
CO
a.
CD

CD
a
<

CD CD co
CI C CO
o o CD O
to C a)
>< s. g
to
3
CQ

2 TJ
ca CO C
O
X CO :
x "CD CO
c o
c D)
ca ca 3
55 CO CQ

o>
o
CM
CD
I

r--
a co
co

CO

o 3
o ca O
XJ
CC o k-
o CO CO
O CD
O
I
CD
TJ
co
CO CD
CQ
D
CD CD CO CO
<
c c c c
2 22 2 CD CD CD ID <D
CD CD CD CD CD CD CO CD CD
C C CZ CZ CL
to to to to Cfl CO Cfl Cfl CO
c c c c O O O TJ TJ TJ TJ c x: x: XI SZ
X o o CO CO CO CO cn
2 2o o >r >> >?
C TJ
TJ C c c
TJ TJ TJ TJ TJ
CO CO 2 2 ttfi C
CO
k-
CD CO CO
D CO
CD CD CD CD W

1 CO CD k_ L_ L L L L L L L L L L L L L L L L L L L L
o
CO
II
o o o o O O CO cO CO co
x: xi C O
CD CO
CD co
XI x:
CO
sz
CO
TJ TJ
co ca
CD
C
CD
fT
CD
.C
CD CD
r. r
CD
x: x: CO CO
CD CD co co
o o o CD CD CD CO
CO
CD
Cfl
CD
CO
CD
CO
CD
CO CO
CD CD
co co xi LL
xi LLx: 55
X
LU
a CD CD
i_ i_
LL
CD LL
CD LL
co co co >< >- 55 55555
LL LL LL CO CD CD
X sz >, >.
i g) g) X sz
X X 2> 2>
p X X
o CM CO
X T- CM co * in co co
$ cn cn cn cn i - C M C o - * m c o i ^ c o o ) ~
LL CO CO CO CO CO
Lu CQ CQ CQ CQ X X X X I X X X X
z Ui Ui Ui Ui Ui ui ui ui cn U) Ui Ui Ui Ui Ui C
^O
. >* in
CM CO * <"
CQ CQ CO CQ CQ co co co co
p
8 CD
OO O) O i- CM n * m co N oo o> O T- CM n ^t in ID s c o o o T - C M c o T t w c o r * -
LU 3
IO CO N N N |^ t^ 1^ N N N S 00 CO CO 00 CO 00 00 00 C O C O O ) 0 ) O ) O > O ) O ) 0 ) 0 )
_l a co co co co co
Q- co co co co co co co co co co co co co co co cocococococococococo
o m m m m m m m m in m in in in in m in in co in m
X X X X X x i nx
L nxi n
xmxi nxi nxm x
i nx
m ixn
2 X X X X X X X X X X X X X X X
I to
O
CD k_

< p p

Q.
CD
E E *

O
3 3 I
co co c
4^ -^
3 k- o
CO CO
O
5* X : CD
CO
-_l CD
D) g
-J
X c co
CD
\7i o CQ

<

CO

x"
X cocococococococo 10 CO CO CO CO
3 3 3 ^ 3 3 3 3 3 3 3 3 3

I
X
LU
CO CO CO CO CO
r- I - r r - I I I
CO CO CO .CO CO CO CO CO
I- h- \~ h- I-

X
o
z
2
T - C M C O T t W C O C O O ) ' - W ^ W C O S C O O ) j- CM co TI- in co h
o l-l-l-l-l-l-l-H O O O O O O O
X

0
3 OO CJ) CM CO
LU CJ) O) O)
CDt"~000)OT-CMCOTt in co i- oo CD o i-
I
g CO CO o o O O O O - ^ - ^ - ^ - T - T -
^^tTj-Tj-Tl-Tj-TtTl-Tt
t" T- T- T- T- ftl W
X inininininininm
Tl" T* Tf Tf Tf Tf Tf Tl" Tt
CO ininmmininioinm
2 +-< XXXXXXXX in m in m m m m
CO
O
xxxxxxxxx X X X X X X xl
S5
= = = = = j=>->.>,>,>,|=^>,^^^|g>,>,>,>,>,^ = =
rt*S = = = --4j'i.-bbiibb = = - = ibbii
S s s S S S - Q x j j o i J B S r t - Q S S S S S ' j Q S x j x j x s S S S
CO 3 3 3 3 3 3 C O c O c f l C O C 0 3 3 C 0 3 3 3 ^ 3 C O C O C O C O C 0 ^ 3 ^
Q T J T J T 3 T J T J T J ^ ^ - - - - g - g - - g - g - g - g ^ - - - - - - g - a t j
CD T " T " T x _" T - O t) O O T) - - TJ TJ T J TJ T J TJ _- _- -
x x xx xx XX X X X I X X X X
Cfl
en
CD
cfl
CJ)
CD
o

CD CD
CL c X)
Q) CD CD
o o E a) CL CL CL
+. CD
2 2
co
TJ
t Ifl
TJ
CD
C
ca
x: to t o CO
CO _
2cfl c o TJ TJ
c cfl CI c _
CO
CO : O
X CO CO CO
CO O O O
CO
o
D)
"CD
CL
o
X)
cfl
r i Ui Ui Ui
cfl CO CO
o oo
3 2 S E
co
o O O O
CQ
CO o
co o

CD

<
X
h-
co
<
CO
o^o oca CJ o C j 0 0u o12o Z~L
o u12 12
tnCO"5nC0c0COC0C0 =
O o
to "cfl g
o

co o o m O m O O O O O
z > O O o CD O O
< O >o>ooooo 5 oo
X
X
< CD a) CD CD CD CD CD
CD CD O
5 k^ 4^ c c c c c CO c
C
0) 0) 0) 2 2
2 CL
X O CO o o o 2 2c o cfl CO cflco ca
CO
o TJ
CO
IT,
CO TJ 4* 4 ^ 4 ^ 4^ o x: x: x:x: jr
LU TJ CO CO
<o co to to o Cfl
CD to x: TJ CI
COCD C >> D O O O O O TJ
Ui TJ TJ
Ui Ui TJ TJ
X
>> >> CO COw TJ
c TJ 0)
CO CO
CI) CI) CI) CD CO c CO
H g co cfl TJ T J TJ TJ s? co c CL CD CD a) CD a) TJ 0)C 01 TJ
O CO
C O - CDcfl CO
k_ uCD CD CD CD
ca o CO r r 1- C o TJ
CO . k_ k_
X CO -.i--. k E ^- CO C D C D C D C D r
TJ 0) HI ? ? 2 c CO
4^ CO Cfl 10 Cfl Cfl CD a)
XI
o d u to r O * O H CO Cfl CO CO O 2 CO CO
Hz Cfl CD CD CD CD CI) 0) a> CO c 0)
TJ TJ J= CO r CO
CO CD CD CD CD O TJ CD 5 555 5 CD cfl 4^
w CO CD
z co > CD -s;
5 55 5
CD c
5 _ tocacfl CDLL 5
2
Q.
CO
5 a> 5 >< >* >> >^ >> >> 5 I 5
TCD
n CD CD CO CD
*. *> -4 -J" -'
CD5
> ^1
LL L L 2C11/D WC D WC>?
D2C D2
M/ W C D ^^ L L ^ CO C O
>, CO CO COCO CO- x:
3 > 2 > T Cfl
P
2 to
2 2
4 ^ 4 ^ 4 ^ 4 ^ 4 ^ ^*
ca to x.C^
ft x:
CD CD aj k_

TJ T J TJCD CO*-.
k_
CD CD >>s
X CJ) CD i_ to to k_ k_ 4g> ^ gj CD
co x: CO o
O)
O O O TJ TJCO LL x 2
i

a = TJ CD i_ k_ CD CD X to CO O O
TJ 2
CD CD2 2
TJ 2
TJ CD Cfl 2 2 2 CD CJ) ID
c CO o CD TJ TJ
O TJ TJ O O
CD O O O CO o
CD

i iUUUWHi CO CO CO CO co co co co
gg
CO
<
Z
O
CD S 00 CO o C M C O T l - m C O I ^ - O O O J O i - C M C O T l - l O C O r ^ O O C D O T - C M C O
3 3
a
CO CO CO Tt T t T j - T j - T j - T t T j - T j - T t k O i n i o i n k n i n i n k O i o i n c o c o c o c o
Tj-Tt-Tt-^TtTfTl-Tl-Tl-Tt-TfTt-Tl-Tt-Tl-Tl-TtTl-Tl-Tl-Tl-Tt
UJ TJ" TJ" Tt Tf
o mininininmininininininLninininininininminmmininin
_l
X
co xxxxxxxxxxxxxxxxxxxxxxxxxxx
I
CO
c
o
TS
CD
CO
c
XI
1-
4^ 4^ 4^

c 10 to to CO CO
CD CD CD CD
o 1- |2 h- h- 1-
cfl " " CO CO
to Cfl CO "
CD CD CD
k_ CO CD
XI XI
co X XI x:
CO CO
a Ui co co
5T
i_
>> > >i >. >> > k. W k_ k- k- l~ k_ l_ k_
X k- k_ k- k- k_ 1
cfl cc cfl
CD CO CO CO
cfl Cfl CO CO Cfl CO CO CO CO
c c c c c c c c c c c
cn c c c kc. k- k_
k. k_ I k_ k_ k- I 1 1_

< co CD a) CD CD a) Ct) CO CD CD CD CD 2 2 2
CO CO cfl co co a CO CO CO CO CO CO CO CO CO
3 3 3 3 3 3 3 3 3 3 3 3 3 3 3

a oaP P Pa p a a a aOOP

c
_g CO CO CO CO CO
3 3 3 3 3
to cfl = = cfl = = CO = = CO = = CO = =
E \- 1- h-
k.
o
X

O) 0) O) O) CJ>
(-~
o o o o o
X CM CM CM CM
CM
TJ CD . . CD .

. CO . . CD .
i
.
CO . .
"i 1
co 1^
CD CO CO
CO CO
CO CO CO CO CO

Q.
< 2 CL
X E a. D.

h-
o 3 E E
X CO 3
co c X CO 3 3
to CO CO
o g
Zi k_ Jki
CO k_ .
< o = = 3 " " o
o
_l
o" X = =
o O CO
5 CD
TJ
CO
CJ) X XJ o
CO c >. cfl O
"cn X
z CD o "5
CO o
< 2
X
X
X
LU
I
X
h-
c
o
a , CO CO CO co CO
X *k_ 3 3 3 3 3
o CO = " CO = = "CO = "
p cn H H h-
p CD
X Q
LL
Cfl

cfl
CD
XI
10 a
Z
TJ
i "s lis 1 g 1 g 1g
CD O CD o 2
o LL
ox o i i- 2 x2 |2 glf I 2
O ^
Q Q>
CM CO * LO CO N- 00 0) o T-
CM CO Tf LO CD
o I "^
_j 3
CM CM CM CM CM CM CM CM CO
LU D)
T* Tf Tf Tl" TJ" Tf
CO
CO CO CO CO CO
o Tf Tfr Tt Tf
Tf Tt TI- TJ- ra-
< cfl m m m m in w m m in in in in in in in
X X X X X X X X X
CO o X X X xxx
X) CO T- CM
O) 00 i- CO TJ-
O T- CM CD CO
n
n
in o
n *
co Tt
^i ^
r ^ S O)

CO 00

*-
Q. co r- r- ^; co
QCD oo 0) CD CM m
0) CO O) TJ- CO
CM CO CO TJ-
co
CD
r h-
o TJ
CO
XJ X J .
CO CO - CO o = = = =
k_ k. CO_l
3 3 Q 4^
CI) c
TJ TJ
X o
Q-
g c .2 .2 c
CO CO CO
"io CO CO CO ~
to .co co E
0) m .*- ir
co CD rr I- = I- = 2
O)
X < Q> k_ X
< CD
CD v % 2
o 4'
CO
1 !z
CD CD CD
c CD S S a)
CD o c O to
2 to O B S:
o
C
o to g 2 co^? J |
TJ TJ g
< to 5 w
X TJ
c
CO
C
Cfl
4-"
CO co I
co 2 c3> co M-
C
h- CO : CO CO E
L>
CO X;= x: t CO
3 XJ
Cfl CO XI : o X S)= 2 E O
=) CJ) -^' CD o O _
o 3 to
< gj O o
2 5 -a > cc
CO
3 k.
O m c i- > o
z CD o
XJ 2 S o
< k_
CO O) CO CO
X o is.
X < t5 o
CO
< X < X CM
CO . . . -
h- _l TJ is.

5 CD co CD < CD co
z
< < X
h-
i
CO
X co
LU 5 zo
<
X CO
h- z 5
CO
X >
<
o X
c c
o o o
z
<
X
" ^

UJ
O
to
X o
zCD <
O O O X
< _J CD
n 5
2c z 5
*.
g X
o
to LU
_i

X
TJ X
z
X
c LU
X X
cfl CD CD f-
CO o c c X
Q.
TJ TJ TJ TJ TJ TJ
CO CD CD CD CD CD
o 2 2 o
3CD i
CO

CO
k_
CD
k-
0)
k_
CD
k_
CD
X to to
z
2
CD
CL
a,
CL
CO
CL
CD
CL
OCD X c g o a) g
CD
XJ
CO
2 2 2 xCO COrCO r
to"
TJ
c c
TJ O o TJ
*-t
0
CO to
>, TJ
CO XI
CO TJ
CO
X
X
5cflCO 5 55
cfl cfl
CD 0)
55
CD CO CO CO CO _
Q.
g C CO C CO 5
3 CO CO cfl o CO o CO -^ CO
CD CD CD CD CD CD CO 3
O CI) XI XL SZ CD XI
(O
CO 4^ 4^ 4^ 4^ 4^ - CO
TJ TJ g Q CO S <" LL ">
Cfl Cfl Cfl cfl CO cfl CD CD
TJ
CD k- k- k- k. L. k_
O c
CD P
LL X
CD "" CD
LL X
XJ eo co co CD CD CD O M CD
k_ TJ T J TJ TJ TJ TJ CO
2 2 c^ XJ
0)
L_ C O
2 O O O
o o o L_ CO CO iL
CO CO
c CO CO
Zk_
CD
C ICD 5 5 CD
o
CJ.
>.
CO CO>. 3
Q. O r
3 x:c xic Z
I 8888 2g> g)
2
Xto Xto
Z
u c
a
c%cl&uiu}
Bft 5 o
2
< < TJ TJ < X
O
Oo o C C CT
CO CO _l a
O CD
^ in
ID N co o) O i- CM
CO CO LU
co Tt in CD h-
_j
LU
2J
CJ) CO CD CD CO CO CO
o N N- N _l
ct rs.
TJ"
i^
TJ"
is.
Tf
r iv.
Tf Tl-
_l Tf TJ- TJ r m m m in in
_l g TJ" Tf TJ" TJ" Tf TJ in in m X
m m m in m m UJ <2 cc{ X
X CO
_l XXX X X X X
to X X X X X X CO C
i O x
CO 2
<
CO
GLOSSARY
GLOSSARY

Arenite: A general n a m e used for consolidated sedimentary rocks composed of sand-


sized fragments irrespective of composition; e.g. sandstone.
Breccia: A coarse-grained clastic rock, composed of granular broken rock fragments
held together by mineral cement or in a fine-grained matrix, it differs from
conglomerate in that the fragment have sharp edges and unworn corners.
Clay gouge: A clayey deposit in a fault zone; fault gouge.
Clay shale: A shale that consists chiefly of clayey material and that becomes clay on
weathering.
Claystone: A n indurated clay having the texture and composition of shale but lacking
its fine lamination or fissility
Fault breccia: A tectonic breccia composed of angular fragments resulting from
crushing, shattering, or shearing of rocks during movement on a fault, from friction
between the walls of the fault, or from distributive ruptures associated with a
major fault; a friction breccia.
Feldspathic: Said of a rock or other mineral aggregate containing feldspar.
Gouge: A thin layer of soft, earthy fault-comminuted rock material along the wall of
a fault or vein or between layers in the country rock.
Litharenite: A sandstone containing more than 2 5 % fine-grained rock fragments, less
than 1 0 % feldspar, and less than 7 5 % quartz, quartzite, and chert in the framework
grains.
Lithic arenite: A sandstone containing abundant quartz, chert, and quartzite, less than
1 0 % argillaceous matrix and more than 1 0 % feldspar, and characterised by an
abundance of unstable materials in which the fine-grained rock fragments content
exceed the amount of feldspar grains.
Matrix: T h e finer-grained material enclosing, or filling the interstices between the
larger grain or particles of a sediment or sedimentary rock.
Mudstone: A n indurated m u d having the texture and composition of shale, but lacking
its fine lamination or fissility; a blocky or massive, fine-grained sedimentary rock
in which the proportions of clay and silt are approximately equal; a non-fissile
m u d shale.
Quartzarenite: A sandstone that composed primarily of quartz; specifically a sandstone
containing more than 9 5 % quartz framework grains (excluding detrital chert
grains).
Quartzose: Containing quartz as a principal constituent.
Shale: A fine-grained detrital sedimentary rock, formed by the consolidation of clay,
silt or m u d . It is characterised by finely laminated structure, which imparts a
fissility approximately parallel to the bedding.
Soil (1): All unconsolidated materials above bedrock. This is c o m m o n usage among
engineering geologists and is the definition adapted in this thesis.
Soil (2): In the engineering sense, any of the drift deposits forming part of the Earth'
crust, except for the agricultural topsoil, which are not part of the solid rock formation.

Вам также может понравиться