mass compressibili'iy of fractured chalk viepubs.surrey.ac.uk/773029/1/359742_vol1.pdf · the...

246
THE MASS COMPRESSIBILI'IY OF FRACTURED CHALK VI - I r \ .j, ,I A thesis submitted to the University of Surrey for the Degree of Doctor of Philosophy in the Department of Civil Engineering by MARCUS CHARLES MATTHEWS B.Sc.(Hons.), M.Sc., F.G.S. Volume 1 Introduction & Literature Review DECEMBER 1993

Upload: duongduong

Post on 14-Jul-2018

214 views

Category:

Documents


0 download

TRANSCRIPT

THE MASS COMPRESSIBILI'IY OF FRACTURED CHALK VI (",.fr-t>~· - ~~ I r \ .j, ,I

A thesis submitted to the University of Surrey for the Degree of

Doctor of Philosophy in the Department of Civil Engineering

by

MARCUS CHARLES MATTHEWS

B.Sc.(Hons.), M.Sc., F.G.S.

Volume 1

Introduction & Literature Review

DECEMBER 1993

To

my Father

Charles Franklin Matthews

SUMMARY

This thesis is concerned with the mass compressibility of fractured chalk and

its influence on the settlement of shallow foundations. A review of the

literature reveals nineteen case records of load-settlement behaviour from

relatively small diameter « 1m) plate loading tests but only six well­

documented case records of the behaviour of shallow foundations on chalk.

The plate loading tests indicate that highly fractured near-surface chalk

undergoes yield at relatively low stresses (200 - 400kPa) resulting in a

significant reduction in stiffness. This behaviour contrasts with that observed

in other rock types with similar discontinuity patterns. For chalk it has only

been observed in one case record for a full-scale foundation. Little is

understood about the mechanisms causing yield.

At the time of starting this research, based largely upon the experience

..gained from in-situ loading tests carried out at Mundford, Norfolk (Ward et

aI., 1968), it was known that factors such as fracture spacing and aperture

played an important role in controlling the load-settlement behaviour of

shallow foundations. Little attention was paid to the large variation in intact

properties displayed by the chalk. In this research nine 1.8m diameter plate

loading tests have been carried out by the author on chalks with different

intact mechanical properties and similar discontinuity patterns. These data

are used to evaluate other in-situ tests (such as SPT, surface-wave geophysics

and visual assessment) as means of providing parameters for the prediction

of foundation settlement.

The results of this research indicate that fractured near-surface chalk

undergoes yield within the range of stresses likely to be imposed by shallow

foundations and that the pre-yield stiffness of the rock mass is controlled to a

large extent by the looseness of the fracture-block system, which in tum

appears to be associated with the intact mechanical properties of the rock.

The post yield-stiffness of the rock mass is generally about one tenth of the

pre-yield stiffness and is relatively insensitive to the rock material properties.

1

ACKNOWLEDGMENTS

The author would like to thank Prof. C.R.!. Clayton for his help and guidance

in carrying out the field work for this project and in the preparation of this

thesis. Special thanks are due to Chris Russell without whose hard work and

expertise the field work could not have been completed.

The author would like to thank the Building Research Establishment for the

loan of the 500 tonne plate loading test apparatus and Tony Butcher for his

help and advice concerning plate loading tests. The writer is particularly

grateful to Tony Butcher and Colin Abbiss of BRE for introducing the author

to surface-wave geophysics, and for their help with the fieldwork at

Leatherhead.

The author is also grateful to the following people for their help with the

preparation of the test sites and assembling the plate testing equipment:

Paul Cheesman

John Feldman

Bob Hillier

Vicki Hope

Roger Hopper

Brian Inch

Colin Sivewright

Ian Wilkinson

Peter Williams

Rick Woods

Special thanks are due to Adriano Bica whose help and advice with the

instrumentation proved invaluable.

11

The following organisations and individuals helped in providing the test sites

and assistance with site preparation:

Esso UK

Higgs and Hill pIc

Needham Chalks Ltd

Strott and Parker

The Expanded Piling Company

Brian Annis

Mr and Mrs Findley

John Seaman

Special thanks are given to Suffolk County Laboratory and Mike Burch for

carrying out Standard Penetration Tests and Dynamic Probing at the

Needham Market test site. Dynamic Probing was also carried out both at

Needham Market and at Leatherhead by Kevin McElmeel of BRE.

The author would like to thank Bill Ward for helpful discussion on the

background to the pioneering work on the mass compressibility of chalk

carried out at Mundford, Norfolk in the 1960's

The author is particularly grateful to Kevin Shaughnessy, Ian Davis and Chris

Bussey for their assistance in the production of this thesis.

Funding for this project was provided by an SERC research grant

(GRID 195883).

Finally the author would like to thank his parents for their support during the

period of this project.

111

CONTENTS

VOLUME 1

1.0 INTRODUCTION

2.0 LITERATURE REVIEW

2.1 Deposition, Diagenesis, Geological Structure

and Weathering

Deposition and Diagenesis

Mechanical Properties of Intact Chalk

Discontinuities in Chalk

Weathering

Summary

2.2 Mass Compressibility

Influence of Discontinuities

Load-Defonnation Behaviour of Fractured

Chalk

Behaviour of Foundations on Chalk

Nonnalised Load-Settlement Behaviour

Plate Loading Tests on Chalk

Summary

2.3 Methods of Assessing the Mass Compressibility

of Chalk

Pressuremeter Tests

Plate Load Tests

Geophysics

The Standard Penetration Test

Vzsual Assessment

Summary

2.4 Summary

. IV

Page

1

5

7

8

13

18

25

31

45

45

67

68

67

89

95

137

140

146

157

166

173

180

199

VOLUME 2

3.0 LOCATION OF TEST SITES AND

CHARACTERISATION OF CHALK

3.1 Test Site A: North Ormsby

Site Location

Topography

Geology

Rock Material and Rock Mass

Characteristics

3.2 Test Site B: Leatherhead

Site Location

Topography

Geology

Rock Material and Rock Mass

Characteristics

3.3 Test Site C: Needham Market

Site Location

Topography

Geology

Rock Material and Rock Mass

Characteristics

v

Page

204

208

208

209

210

215

235

235

235

236

237

260

260

261

261

264

Page

4.0 FIELD AND LABORATORY TESTING TECHNIQUES 287

4.1 Plate Loading Tests 290

Apparatus 292

Instrumentation 296

Procedure 311

Results of Plate Loading Tests 319

4.2 Other In-Situ Tests 380

Standard Penetration Tests 380

Dynamic Probing Tests 387

Geophysics 391

4.3 Laboratory Tests 427

Dry Density 428

Uniaxial Compressive Strength 433

Brazilian Tensile Strength 436

Intact Stiffness Characteristics of Chalk 437

On-Dimensional Compression Tests

on Intact Chalk 440

VI

VOLUME 3

5.0 DISCUSSION 453

5.1 Mechanical Properties of Intact Chalk 460

Dry Density 460

Uniaxial Compressive Strength 461

Brazilian Tensile Strength 466

Stiffness in Uniaxial Compression 468

Load-Deformation Behaviour

in Uniaxial Strain 471

Summary 476

5.2 Rock Mass Description 499

5.3 Load-Settlement Behaviour of Chalk 513

Initial Modulus E1 515

Post-Yield Modulus E~ 520

Yield Bearing Pressure q£. 522

Yield Bearing Pressure q~ 523

Collapse Settlement at

Needham Market (Site C) 524

Creep 526

Summary 531

.. Vll

5.4 Methods of Determining Mass Stiffness Parameters

Settlement Predictions using

Surface-Wave Geophysics

Settlement Predictions using

the Standard Penetration Test

Settlement Predictions using

Visual Assessment of the Rock Mass

Summary of Settlement Predictions

Cost Effectiveness of the Methods

of Settlement Prediction

Summary

5.5 Design of Plate Loading Tests

5.6 Design of Spread Foundations on Chalk

6.0 CONCLUSIONS AND RECOMMENDATIONS

6.1 Conclusion from Literature Review

6.2 Conclusions from Laboratory Studies

6.3 Conclusions from Field Studies

6.4 Implications for the Design of

Spread Foundations

6.5 Recommendations for Further Work

REFERENCES

APPENDIX A

APPENDIX B

Drill-hole logs, trial pit logs and face logs

Creep rate data

Vlll

Page

543

544

548

557

561

564

568

579

589

595

597

602

604

612

613

615

639

662

LIST OF TABLES

Table Title Page

Chapter 2

2.1/1 Zonal division of the Upper Cretaceous. 9

2.1/2 Range of index properties for the Chalk. 14

2.2/1 Characteristic load-deformation 57 behaviour for rock masses. (After Barton, 1986).

2.2/2 Summary of published case records of 69 the behaviour of foundations on chalk.

2.2/3 Parameters derived from average surface 72 settlement measurements (after Kee, 1974).

2.2/4 Percentage of immediate strains 77 recovered on unloading and values of Young's Modulus (after Ward et al. 1968).

2.2/5 Plate loading tests on chalk. 90

2.3/1 Methods used to determine the 139 deformation parameters of chalk.

2.3/2 Values of Poisson's Ratio determined 151 from unconfined compression tests (from Bell et al., 1990).

2.3/3 Principal types of elastic wave. 159

2.3/4 Basis of Wakeling's tentative correlation 168 between SPT 'N' and elastic modulus (from Wakeling, 1966).

2.3/5 Extended visual classification of the 176 chalk (after Clayton, 1990).

2.3/6 Correlation of engineering grade with 178 mechanical properties of chalk in the mass.

IX

Table Title Page

Chapter 3

3.1/1 Classification of the Upper Cretaceous 213 in the Northern (Yorkshire and Lincolnshire) and Southern (southern England) provinces (after Kent and Gaunt, 1980).

3.1/2 Field description of hardness. 216

3.1/3 Details of vertical joint sets at North 217 Ormsby.

3.3/1 Sub-vertical joint spacings. 265

Chapter 4

4.1/1 Rock Mass compressibility parameters 321 derived from plate loading tests.

4.2/1 Summary of SPT tests carried out at 381 plate test sites.

4.2/2 Summary of SPT tests carried out at site 382 A (North Ormsby).

4.2/3 Summary of SPT tests carried out at site 383 B (Leatherhead).

4.2/4 Summary of SPT tests carried out at site 386 C (Needham Market).

4.2/5 Test parameters for ASTM 1586 and 389 DIN 4094.

4.2/6 Details of Surface-Wave Tests at site A 401 (North Ormsby).

4.2/7 Details of Surface-Wave Tests at site B 403 (Leatberhead).

4.2/8 Details of Surface-Wave Tests at site C 404 (Needham Market).

4.3/1 Summary of dry density measurements. 432

x

Table Title Page

4.3/2 Summary of unconfined compressive 435 strength test results.

4.3/3 Summary of brazilian tensile strength 436 test results.

4.3/4 Summary of stiffness measurements on 440 intact chalk.

Chapter 5

5.1/1 Summary of uniaxial strain test results. 474

5.2/1 Assessment of looseness of the fracture- 503 block system.

5.3/1 Summary of rock mass compressibility 515 parameters derived from plate loading tests at the three test sites.

5.3/2 Rock mass factors for the three test 520 sites.

5.3/3 Ratio of qe to intact yield stress «(Jy). 523

5.4/1 Values of Go and m derived from 547 surface-wave seismic tests.

5.4/2 Ratio of observed settlement and 548 predicted settlement based on stiffness profiles determined from surface-wave seismic tests.

5.4/3 SPT N values. 553

5.4/4 Ratios of observed to predicted 554 settlements for a bearing pressure of 200kPa based on the Standard Penetration Test.

5.4/5 Degree of loading ('bet/ q(ult)u) for a 556 bearing p~essure of 200kPa .

. Xl

Table Title Page

5.4/6 Results of settlement predictions for a 562 bearing pressure of 200kPa based on visual assessment using the Mundford grading scheme devised by Ward et al. (1968).

5.4/7 Assessment of mass compressibility of 565 chalk based on visual assessment.

5.4/8 Logistics of in-situ measurements of 567 stiffness.

5.4/9 Unit cost of in-situ stiffness 568 measurements.

XlI

LIST OF FIGURES

Figure Title Page

Chapter 2

2.1/1 Variability of the chalk in terms of dry 33 density and porosity (from Clayton and Matthews, 1987).

2.1/2 Results of density tests on chalk from 33 West Surrey, Isle of Wight and the Isle of Purbeck (after Clayton and Matthews, 1987).

2.1/3 Variation of uniaxial compressive 34 strength of chalk with dry density and porosity (Data from Masson, 1973, Bell, 1977, Bonvallet, 1979, Woodland et aI., 1988, Clayton and Safari-Shooshtari, 1990, Bell et aI., 1990, Blight, 1990, Kroniger, 1990, Mortimore and Fielding, 1990, Nienhuis and Price, 1990, and Varley, 1990).

2.1/4 Variation of Brazilian tensile strength of 34 chalk with dry density and porosity (Data from Bell, 1977, Bell et aI., 1990, Kroniger, 1990, Mortimore and Fielding, 1990 and Nienhuis and Price, 1990).

2.1/5 Relationship between the uniaxial 35 compressive strength of dry and saturated specimens of chalk (Data from Masson, 1973, Bell, 1977, Bonvallet, 1979, Bell et al., 1990, Blight, 1990 and Kroniger, 1990).

2.1/6 Variation in stiffness of chalk with dry 35 density and porosity (Data from Bell, 1977, Bonvallet, 1979, Woodland et al., 1988, Bell et al., 1990, Kroniger, 1990 and Nienhuis and Price, 1990).

2.1/7 Typical stress-strain behaviour for intact 36 chalk (after Jardine et al. 1985).

Xl11

Figure Title Page

2.1/8 Variation of secant Young's modulus 36 with axial strain for chalk (after Jardine et al. 1985).

2.1/9 Yield observed in chalks of different 37 porosities when subjected to uniaxial (~) compression (after Leddra et al. 1993).

2.1/10 Map of southeastern England showing 37 the main structures affecting the Chalk.

2.1/11 (a) Block diagram showing relationships 38 between effective principal stresses and an extension fracture (E) and conjugate shear fractures (S) developed in a mechanically isotropic brittle rock. Stipple indicates the quadrants within which hybrid fractures form. (b) Composite failure envelope and Mohr circles constructed for 2a = 0, 45 and 60°. (T = tensile strength and ¢ = angle of internal friction. (after Hancock, 1985).

2.1/12 Rose diagrams showing typical grid lock 39 and joint spectrum discontinuity patterns in rock (after Rawnsley et al. 1990).

2.1/13 Typical system of jointing developed on 39 the linb of a fold (after Price, 1966).

2.1/14 (a) Frequency polygons for dominant 40 joint sets plotted against their strike for the Chalk recognised in the Chilterns. (b) Map of strikes of fracture sets observed in the northern Chilterns. (after Cawsey, 1977).

2.1/15 Frequency polygons showing the 41 variation in dip of the major joint sets recogniseg in the Chilterns (after Cawsey, 1977).

XlV

Figure Title Page

2.1/16 Block diagram illustrating orientation 42 and frequency of joints and faults in the chalk near Culver Cliff, Isle of Wight (after Fookes and Horswill, 1970).

2.1/17 Typical weathering profile for the Chalk. 43

2.1/18 Variation of vertical joint spacing with 44 depth for four sites in East Sussex (after Williams, 1987).

2.1/19 Variation of sub-horizontal fracture 44 spacing with depth for chalk, based on drillhole logs for Princes Quay, Hull (after Woodland et al. 1988).

2.2/1 Concave normal stress-deformation 99 behaviour of rock joints (after Bandis et al. 1983).

2.2/2 Measurements of the closure of natural 100 joints under normal stress (after Bandis et al. 1983).

2.2/3 Convex shear stress-displacement 100 behaviour of rock joints illustrating sample size dependence (after Bandis et al. 1981).

2.2/4 Mated and non-mated joints. 101

2.2/5 Relationship between joint contact area 101 and average normal stress for a variety of rock types including chalk (after Duncan and Hancock, 1966).

2.2/6 Relationship between joint contact area 102 and average normal stress for natural fractures in quartz monzonite (after Pyrak-Nolte et al. 1987, 1990).

xv

Figure Title Page

2.2/7 Constrasting load-deformation behaviour 103 for rock masses with different magnitudes of joint shear (S) and normal deformation (N) components (after Barton, 1986).

2.2/8 (a) Deformational response where the 104 joint sets are perpendicular to the imposed principal loads (after Chappell, 1979). (b) Results of a plate loading test on sandstone with predominantly sub-horizontal and sub-vertical joint sets (after Hobbs, 1973).

2.2/9 Simple models for studying the influence 105 of joint spacing on mass compressibility.

2.2/10 Variation in the ratio of mass stiffness to 106 intact stiffness with fracture frequency for fractures with different stiffnesses.

2.2/11 Variation in the ratio of mass stiffness to 106 intact stiffness with fracture frequency for fractures with different apertures.

2.2/12 Variation of rock mass factor (j) with 107 fracture frequency for chalk based on laboratory tests on artificial joints (after Wakeling, 1975).

2.2/13 Variation of rock mass factor (j) with 107 fracture frequency for joints with different initial contact area ratios.

2.2/14 Variation of rock mass factor (j) with 108 joint roughness expressed as the number of asperities per unit area (Na) for rock masses with different joint spacings.

2.2/15 Variation of rock mass factor (j) and 108 fracture frequency for chalk (after Hobbs, 1975).

XVI

Figure Title Page

2.2/16 Pressure distributions for a circular 109 foundation on a rock masses with different discontinuity set orientations (after Gaziev and Erlikhman, 1971).

2.2/17 Typical pressure-settlement curve for 110 Grades IV and III chalk (from Burland and Lord, 1970).

2.2/18 Silo foundations and positions of 111 levelling points and inspection shafts (after Burland and Bayliss, 1990).

2.2/19 Chalk grades based on visual inspection 112 of the chalk beneath the silos in shafts 1, 2, 3 and 4 (after Kee, 1974).

2.2/20 Settlement of Silo No.3 during first 113 loading (after Burland and Bayliss, 1990).

2.2/21 Settlement profiles for silo No. 3 during 113 first loading (after Burland and Bayliss, 1990).

2.2/22 Distribution of settlement with depth 114 beneath the centre of silo No.3 for various foundation pressures (after Burland and Bayliss, 1990).

2.2/23 Vertical strains beneath the centre of 114 silo No. 3 (after Burland and Bayliss, 1990).

2.2/24 Relationship between vertical stress and 115 depth beneath centre of silo when filled to capacity (after Nicoletto, 1979).

2.2/25 Stress distribution beneath a silo when 115 filled to capacity (after Nicoletto, 1979).

2.2/26 Vertical section showing the position of 116 the instrumentation used to monitor ground movements associated with the tank loading test at Mundford, Norfolk (after Ward et ale 1968).

XVII

Figure Title Page Section beneath tank showing the 117

2.2/27 distribution with depth of the immediate vertical deflections under maximum tank load at shafts 1, 2, 3 and 4, and the deflected shape of the ground surface (after Ward et aI., 1968).

2.2/28 Relationship between pressure and 118 immediate deflection at various levels in shafts 1 and 2 (after Ward et aI., 1968).

2.2/29 Short-term tank test; relationship 119 between time and vertical strain at various levels beneath the centre of the tank during loading, standing and then unloading (after Ward et aI., 1968).

2.2/30 Long-term tank test; relationship 119 between vertical deflection of level 1 (relative to level 6) in each shaft after completion of loading. The differential settlements at level 1 between the centre (Sl) and the edge (S2) of the tank are also shown (after Ward et aI., 1968).

2.2/31 Long-term tank test; relationship 120 between time and vertical strain for three levels beneath the tank over a period of one year (after Burland, 1975).

2.2/32 Profile of chalk grade at Luton, site D 121 (after Powell et aI., 1990).

2.2/33 Pressure-settlement curves for the slab 122 loading test carried out at Luton, site D (after Powell et aI., 1990).

2.2/34 Pressure-settlement curve for the slab 122 loading test at Luton, site D with the creep settlement removed (from Powell et al., 1990).

2.2/35 Profile of chalk grade at Luton, site C 123 (after Powell et al., 1990).

XVlll

Figure Title Page

2.2/36 Pressure-settlement curves for the 124 1710mm pad loading test carried out at Luton, site C (after Powell et aI., 1990).

2.2/37 Pressure-settlement curve for the 124 1710mm pad loading test at Luton, site C with the creep settlement removed (from Powell et al., 1990).

2.2/38 Simplified plan showing the arrangement 125 of the 1.98m and 3.35m square footings at Basingstoke (after Lake and Simons, 1975).

2.2/39 Relationship between time and average 125 settlement for the 1.98m and 3.35m square footings (after Lake and Simons, 1975).

2.2/40 Plan of raft foundations showing the 126 position of levelling stations boreholes and trial pits (after Burland et aI., 1975).

2.2/41 Profile of chalk grade based on SPT 'N' 126 values (after Burland et aI., 1975).

2.2/42 Relationship between time and 127 settlement for some of the levelling stations in Block A (after Burland et aI., 1975).

2.2/43 Cross-section through load test in the 128 base of the coffer-dam showing concrete cylinder, deep settlement points and kentledge (after Burland et aI., 1983).

2.2/44 Results of loading test in base of coffer- 129 dam (after Burland et aI., 1983).

2.2/45 Time-settlement observations on cable 129 chamber (after Burland et al., 1983) .

. XIX

Figure Title Page

2.2/46 Relationship between applied stress and 130 settlement-ratio showing pre-yield and post-yield behaviour for full-scale structures and large-scale loading tests.

2.2/47 Pressure-settlement curves for plates of 131 various diameters (after Lake and Simons, 1975).

2.2/48 Pressure-settlement curves for plates of 132 various diameters (after Hodges, 1976).

2.2/49 Relationship between modulus and plate 133 diameter of widely jointed Maarstichtian chalk (Nienhuis and Price, 1990).

2.2/50 Relationship between load-intensity and 134 creep ratio R from plate tests (after Burland and Lord, 1970).

2.2/51 ( a) Relationship between creep rate and 135 time from plate test at Luton, Site D (after Powell. 1990).

2.2/51 (b) Relationship between creep rate and 136 time from plate test at Luton, Site C (after Powell. 1990).

2.3/1 Typical arrangement for a Menard 185 pressuremeter test.

2.3/2 Typical curve from a Menard 186 pressure meter test in moderately weak rock (after Ervin et aI., 1980).

2.3/3 Moduli determined from pressuremeter 186 and plate tests on chalk compared with those back-analysed from a bridge foundation (from Marsland and Powell, 1983 and Powell et aI., 1990).

2.3/4 865mm diameter loading plate with four- 187 point sub-plate ground-deformation measuring system (after Marsland and Eason, 1973).

xx

Figure Title Page

2.3/5 Typical load-settlement curves from sub- 188 plate deformation measurements (after Hird et aI., 1991).

2.3/6 Error in modulus values derived from 188 sub-plate deformation measurements (after Hillier, 1991).

2.3/7 Typical relationship between stiffness 189 and strain for soils.

2.3/8 Seismic methods for determining the 190 variation of stifness with depth in soils and rocks.

2.3/9 Velocity profile along seismic line 7 at 191 Mundford, Norfolk (after Grainger et al. 1973).

2.3/10 Typical time-distance graph from a 192 seismic refraction test at Mundford, Norfolk (after Abbiss, 1979).

2.3/11 Relationship between static and dynamic 192 moduli for the chalk at Mundford, Norfolk (after Abiss, 1979).

2.3/12 Relationship between Young's modulus 193 and depth for the Mundford chalk determined using seismic refraction, finite element analysis and plate loading tests (after Abbiss, 1979).

2.3/13 Shear modulus-depth profiles for chalk 193 measured using cross-hole and down-hole seiesmic tests (after Sigismond et aI., 1983).

2.3/14 Correlation of stiffness with standard 194 penetration test results for chalk (after Wakeling, 1966).

2.3/15 Correlation of stiffness with standard 194 penetration test results for chalk (after Wakeling, 1970).

XXI

Figure Title Page

2.3/16 Comparison of empirical correlations 195 between E and SPT 'N' value for the chalk.

2.3/17 Variation of E' /N60 with degree of 196 loading for chalk (after Stroud, 1988).

2.3/18 Comparison of the results of standard 196 penetration tests carried out by different contractors at the same chalk site (after Mallard, 1977).

2.3/19 Reported ranges of moduli and yield 197 bearing pressure for in-situ chalk under foundation loading (after Clayton, 1990a).

2.3/20 Results of Standard Penetration Tests 198 and visual grading of the chalk (after Burland and Bayliss, 1990).

Chapter 3

3.0/1 Location of test sites. 207

3.1/1 Location of test site A (North Ormsby). 221

3.1/2 Plan of North Ormsby Quarry. 222

3.1/3 Topography of the area around North 223 Ormsby.

3.1/4 Geology of the area around North 224 Ormsby.

3.1/5 Structure contours for the base of the 225 Chalk in the Humberside area (after Berridge et aI., 1990).

3.1/6 Principal formations for the Chalk of the 226 N orthem Province (after Wood and Smith, 1978).

3.1/7 Corellation of major flint bands across 227 test site A

XX1l

Figure Title Page

3.1/8 Profiles of Mundford grade for site A 228 based on initial on-site appraisal.

3.1/9 Rock mass assessment for face log 1. 229

3.1/10 Rock mass assessment for face log 2. 230

3.1/11 Profiles of Mundford grade for site A 231 based on detailed analysis of spacing, aperture and infill.

3.2/1 Location of test site B (Leatherhead). 242

3.2/2 Plan of the Esso site prior to 243 development.

3.2/3 Topography of the area around 244 Leatherhead.

3.2/4 Geology of the area around 245 Leatherhead.

3.2/5 Limits of the chalk outcrop in western 246 Surrey.

3.2/6 Plan of Esso site showing site 247 invstigation boreholes positions, trial pit locations.

3.2/7 Plan of test site B showing the trial pit 248 and plate test locations.

3.2/8 Weathering profiles seen in deep trial 249 pits at site B. The zones A, B, C and D refer to the weathering zones defined in chapter 2.

3.2/9 Sketch of typical bedding plane 250 discontinuity seen in trial pits at site B.

3.2/10 Profiles of Mundford grade for site B 251 based on initial on-site appraisal of trial pits and foundation excavation.

3.2/11 Rock mass assessment for trial pit TPS 252 1.

XXlll

Figure Title Page

3.2/12 Rock mass assessment for trial pit TPS 253 2.

3.2/13 Rock mass assessment for trial pit TPS 254 3.

3.2/14 Rock mass assessment for foundation 255 excavation.

3.2/15 Profiles of Mundford grade for site B 256 based on detailed analysis of spacing, aperture and infill.

3.3/1 Location of test site C (Needham 269 Market).

3.3/2 Plan of Needham Chalks Ltd Quarry. 270

3.3/3 Topography of the area around 271 Needham Market.

3.3/4 Solid geology of the area around 272 Ipswich.

3.3/5 Superficial deposits in the area around 273 Needham Market.

3.3/6 Zonal map of the Upper Chalk of 274 Suffolk (after lukes-Browne and Hill, 1904).

3.3/7 Plan of test site C showing the trial pit 275 and plate test locations.

3.3/8 Lower hemispherical projection showing 276 the orientation of sub-vertical discontinuities observed in the trials pits at site C.

3.3/9 Correlation of major bedding plane 277 discontinuities observed in the trial pits at site C.

3.3/10 Typical block sizes observed in the trial 278 pits at site C.

XXIV

Figure Title Page

3.3/11 Profiles of Mundford grade for site C 279 based on initial on-site appraisal of trial pits.

3.3/12 Rock mass assessment for trial pit TPS 280 1.

3.3/13 Rock mass assessment for trial pit TPS 281 2.

3.3/14 Rock mass assessment for trial pit TPS 282 3.

3.3/15 Profiles of Mundford grade for site C 283 based on detailed analysis of spacing, aperture and infill.

Chapter 4

4.1/1 Schematic diagram of 500 tonne capacity 330 loading frame used for plate load tests.

4.1/2 Details of loading column used with the 331 500 tonne capacity loading frame shown in Fig. 4.1/1.

4.1/3 Typical tension pile arrangement and 332 attachment to spreader beam.

4.1/4 Typical disc spring and stacking 333 arrangements for disc springs.

4.1/5 Disc spring performance for different 334 stacking arrangements.

4.1/6 Typical arrangement of datum posts and 335 supports for dial gauges.

4.1/7 Typical settlement measurement 336 positions for dial gauges on the 1800mm dia. plate.

4.1/8 Typical bqoking form for dial gauge 337 readings.

xxv

Figure Title Page

4.1/9 Typical output from the spreadsheet 338 used to process the dial gauge readings.

4.1/10 Typical set of processed dial gauge 339 readings displayed as a surface in order to check for errors in data entry.

4.1/11 Typical time-settlement curve derived 339 from the dial gauge readings.

4.1/12 The Zeiss, Jena Ni 007 precise level. 340

4.1/13 View of staff through the Zeiss, Jena Ni 341 007 optics.

4.1/14 Typical arrangement of tension piles 342 used at each test site.

4.1/15 Arrangement of galvanised dome 343 headed levelling stations grouted into 1800mm dia plate.

4.1/16 Typical arrangement of levelling stations 344 used during a plate loading test.

4.1/17 Typical booking form used for precise 345 levelling.

4.1/18 Typical output of spreadsheet used to 346 process precise levelling data.

4.1/19 Typical variation in the apparent 347 movement of temporary bench marks with time.

4.1/20 Schematic diagram of the cantilever 348 beam type load cell used to measure the vertical load on the plate.

4.1/21 Typical calibration curve for the 500 349 tonne Macklow-Smith load cell.

4.1/22 Typical error analysis for load cell 350 calibration.

XXVI

Figure Title Page

4.1/23 Typical sequence of activities associated 351 with carrying out a suite of plate loading tests.

4.1/24 Pressure-settlement relationships from the three plate loading tests carried out

352

at site A.

4.1/25 Pressure-settlement relationships from 353 the three plate loading tests carried out at site B.

4.1/26 Pressure-settlement relationships from 354 the three plate loading tests carried out at site C

4.1/27 Pressure-settlement relationships for 355 tests 2 and 3 at site Coo

4.1/28 Typical time-settlement relationship 356 observed tests 2 and 3 at site which exhibited collapse.

4.1/29 Relationship between secant modulus 357 and bearing pressure for the plate loading tests carried out at site A.

4.1/30 Relationship between secant modulus 358 and bearing pressure for the plate loading tests carried out at site B.

4.1/31 Relationship between secant modulus 359 and bearing pressure for the plate loading tests carried out at site C.

4.1/32 Relationship between tangent modulus 360 and bearing pressure for the plate loading tests carried out at site A.

4.1/33 Relationship between tangent modulus 361 and bearing pressure for the plate loading tests carried out at site B.

4.1/34 Relationship between tangent modulus 362 and bearing pressure for the plate loading tests carried out at site C.

XXVII

Figure Title Page

4.1/35 Typical pressure-settlement curves 363 derived from precise levelling and dial gauges for site A

4.1/36 Typical pressure-settlement curves 364 derived from precise levelling and dial gauges for site B.

4.1/37 Pressure-settlement curves derived from 365 precise levelling and dial gauges for test 1 at site C.

4.1/38 Pressure-settlement curves derived from 366 precise levelling and dial gauges for test 2 at site C.

4.1/39 Relationship between the difference in 367 settlement derived from precise levelling and dial gauges and bearing pressure for all test sites.

4.1/40 Typical pressure-settlement curves for 368 individual levelling stations on the plate at site A showing magnitude of differential settlement.

4.1/41 Typical pressure-settlement curves for 369 individual levelling stations on the plate at site B showing magnitude of differential settlement.

4.1/42 Pressure-settlement curves for individual 370 levelling stations on the plate for test 2 at site C showing magnitude of differential settlement.

4.1/43 Pressure-settlement curves for individual 371 levelling stations on the plate for test 3 at site C showing magnitude of differential settlement.

4.1/44 Relationship between bearing pressure 372 and creep ratio R.

4.1/45 Typical creep behaviour observed 373 during plate tests at site A

XXVlll

Figure Title Page

4.1/46 Typical creep behaviour observed 374 during plate tests at site B.

4.1/47 Typical creep behaviour observed 375 during plate tests at site C.

4.1/48 Time-settlement relationship for the long 376 term maintained load test carried out at site A

4.1/49 Creep rate-time relationship for the long 377 term maintained load test carried out at site A

4.2/1 Plan of site A showing the position of 407 boreholes used for Standard Penetration Tests.

4.2/2 Typical relationship between penetration 408 per blow and penetration measured during Standard Penetration Tests carried out in boreholes advanced using different drilling methods at site A.

4.2/3 Variation of SPT 'N' value and Grade 409 with depth for site A.

4.2/4 Plan of site B showing the position of 410 exploratory boreholes.

4.2/5 Plan of site B showing the position of 411 boreholes used for Standard Penetration Tests and the location of dynamic probe tests.

4.2/6 Variation of SPT 'N' value with depth derived from the site investigation for

412

Esso's new European RO.

4.2/7 Variation of SPT 'N' value and Grade 413 with depth for site B .

. XX1X

Figure

4.2/8

4.2/9

4.2/10

4.2/11

4.2/12

4.2/13

4.2/14

4.2/15

4.2/16

4.2/17

4.2/18

4.3/1

4.3/2

Title

Plan of site C showing the position of boreholes used for Standard Penetration Tests and the location of dynamic probe tests.

Variation of SPT 'N' value and Grade with depth for site C.

Sacrificial cone used for dynamic probing tests.

Typical results of dynamic probing tests carried out at sites Band C.

The principles of continuous surface-wave seismic tests.

Typical output from the spreadsheet used to process continuous surface-wave seismic test data.

Typical graphical output used in the processing of continuous surface-wave seismic test data.

Typical relationships between stiffness and depth from continuous surface-wave seismic tests.

Profiles of shear modulus with depth for plate locations 2 and 3 at site A.

Profiles of shear modulus with depth for plate locations 1, 2 and 3 at site B.

ProfIles of shear modulus with depth for plate locations 1, 2 and 3 at site C.

Variation of dry density and porosity of chalk with depth for each test site.

Variation of uniaxial compressive strength with dry density and porosity of chalk from each test site.

xxx

Page

414

415

416

417

418

419

420

421

422

423

424

443

443

Figure Title Page

4.3/3 Variation of Brazilian tensile strength 444 with dry density and porosity of chalk from each test site.

4.3/4 Typical stress-strain curves for dry chalk 445 tested in uniaxial compression.

4.3/5 Stiffness of dry chalk during shear in 446 uniaxial compression.

4.3/6 Typical stress-strain and stiffness-strain 447 relationships for saturated chalk from site A tested in uniaxial compression.

4.3/7 Typical stress-strain and stiffness-strain 448 relationships for saturated chalk from site B tested in uniaxial compression.

4.3/8 Typical stress-strain and stiffness-strain 449 relationships for dry chalk from site C tested in uniaxial compression.

4.3/9 Variation in stiffness with dry density 450 and porosity of chalk from each test site.

4.3/10 Results of one-dimensional compression 451 tests on chalk from each test site.

4.3/11 Typical relationships between void ratio 452 and vertical stress for chalk tested in one-dimensional compression.

Chapter 5

5.1/1 Range of dry density and porosity found 478 at sites A, Band C in relation to the overall range according to biozone.

5.1/2 Variation of uniaxial compressive 479 strength of chalk with dry density and porosity (data from chapter 2 combined with the results given in Chapter 4).

XXXI

Figure Title Page

5.1/3 Relationship between the uniaxial 480 compressive strength of dry and saturated specimens of chalk.

5.1/4 Variation of Brazilian tensile strength of 481 chalk with dry density and porosity (data from chapter 2 combined with the results given in Chapter 4).

5.1/5 Relationship between uniaxial 482 compressive strength and Brazilian tensile strength of chalk.

5.1/6 Stiffness of undisturbed London Clay 483 during undrained shearing in triaxial compression.

5.1/7 Relationship between secant modulus at 484 0.01 % axial strain and that at 0.001% axial strain for specimens of chalk tested in uniaxial compression.

5.1/8 Relationship between L (EO•01 /EO.OO1) 485 and dry density for specimens of chalk tested in uniaxial compression.

5.1/9 Normalised stiffness of chalk (E/oc) and 486 London clay (E/2Su) during shear.

5.1/10 Variation in stiffness of chalk with dry 487 density and porosity (data from chapter 2 combined with the results given in Chapter 4).

5.1/11 Variation in 'corrected' stiffness of chalk 488 with dry density and porosity.

5.1/12 Relationship between the stiffness of dry 489 and saturated specimens of chalk.

5.1/13 The comparison of structured and 490 destructured compression in the oedometer test (after Leroueil and Vaughan, 1990)

XXXII

Figure Title Page

5.1/14 Yield locus for chalk in void ratio- 491 effective stress space.

5.1/15 Relationship between void ratio and 492 vertical stress for specimens of chalk with different porosities.

5.1/16 (a) Isotropic and (b) one-dimensional 493 compression data for a carbonate sand (after Coop, 1990)

5.1/17 Variation in constrained modulus (l/Illv) 494 of chalk with dry density and porosity.

5.1/18 Comparison between the constrained 495 modulus of chalk and Young's modulus measured in uniaxial compression .

5.1/19 Relationship between yield stress and 496 dry density for specimens of chalk.

5.1/20 Relationship between initial void ratio 497 and yield stress for chalk ..

5.1/21 Variation of compression index Cc of 498 chalk with dry density and porosity.

5.2/1 Principal factors influencing chalk mass 510 compressibility.

5.2/2 Variation of horizontal discontinuity 511 spacing with depth observed at test sites A, Band C.

5.2/3 Principal factors influencing the mass compressibility of chalk which can be

512

assessed by visual inspection of the rock mass.

5.3/1 Characteristic relationships between 536 secant modulus and bearing pressure observed at sites A, Band C.

5.3/2 Pressure-settlement ratio results showing 537 pre-yield behaviour for structures and 1.8m dia. plate loading tests on chalk.

XXXlll

Figure Title Page

5.3/3 Pressure-settlement ratio results showing 538 pre-yield behaviour of structures and 1.8m dia. plate loading tests on chalk classified according to dry density.

5.3/4 Relationship between rock mass factor j 539 and fracture spacing for the chalk (from Hobbs, 1975).

5.3/5 Pressure-settlement ratio results showing 540 yielding behaviour for structures and 1.8m dia. plate loading tests on chalk.

5.3/6 Relationship between bearing pressure 541 and creep ratio R for the chalk based on published data and 1.8m dia. plate loading tests at site A, Band C.

5.3/7 Characteristic relationships between 541 creep rate and log time observed at sites A, Band C.

5.3/8 Relationship between creep rate at 24 542 hours and bearing pressure.

5.3/9 Characteristic relationship between log 542 creep rate and log time observed at sites a, Band C.

5.4/1 Ratio of observed to predicted 571 settlement for bearing pressures of 100, 200 and 300kPa based on continuous surface-wave seismic tests.

5.4/2 Relationship between predicted 572 settlement (normalised by the settlement assuming a Poission's of 0.5) and Poisson's ratio derived from finite element analysis using CRISP90.

5.4/3 Ratio of observed to predicted 573 settlement for a bearing pressure of 200kPa based on empirical relationships between E and SPT 'N' value.

XXXIV

Figure Title Page

5.4/4 Empirical and observed relationships 574 between E and SPT 'N' value.

5.4/5 Visual classification of the chalk at each 575 test site using the Grading scheme for Mundford (Ward et aI., 1968).

5.4/6 Ratio of observed to predicted 576 settlement for a bearing pressure of 200kPa based on visual assessment of the rock mass.

5.4/7 Summary of settlement predictions 577 derived from geophysics, SPT and visual assessment.

5.4/8 Classification of the chalk at sites A, B 578 and C using the scheme proposed in Table 5.4/7.

xxxv

LIST OF PLATES

Plate Title Page

Chapter 3

3.1/1 (a) View of test site A looking south 232 west. (b) View of test site looking north east.

3.1/2 View of quarry face at the location of 233 face log 2.

3.1/3 View of plate location 1 (PL1) during 234 preparation showing loose fracture-block system.

3.2/1 Typical weathering profile seen in the 257 excavation for the foundations of Esso's new European Headquarters at Leatherhead.

3.2/2 View of the transition from structureless 257 chalk into structured chalk in trial pit TPS1 at site B.

3.2/3 Contaminated chalk seen in the 258 excavations for the foundations of Esso's new European Headquarters at Leatherhead. Note how the contaminant (brown discolouration) has picked out the fracture pattern.

3.2/4 Typical open sub-horizontal 259 discontinuities seen in trial pits at site B.

3.2/5 View of slickensides on sub-vertical 259 discontinuity seen in trial pit TPS3 at site B.

3.3/1 ( a) View of test site C looking west. (b) 284 View of test site C looking east showing the arrangement of tension piles and test locations.

3.3/2 View of plate test location 3 during 285 preparation. Note the open sub-vertical discontinuities.

XXXVI

Plate

3.3/3

Chapter 4

4.1/1

4.1/2

4.1/3

4.1/4

4.2/1

4.2/2

Title

View of rock face adjacent to the test site. Note the absence of flaggy chalk near the original ground surface.

View of 500 tonne capacity loading frame during assembly at test site B.

View of loading column showing the load cell, locking ring hydraulic jack and pump unit and the disc spring unit.

Typical arrangement of dial gauges used to measure plate settlement.

View of crushed chalk seen beneath plate location 3 (PL3) at site C. The orange lines show the excavation for the plate loading test and the limits of the crushed material.

Apparatus used to perform dynamic probing tests at sites Band C.

View of the equipment used to perform continuous surface-wave seismic tests at site A

XXXVII

Page

286

378

378

379

379

425

426

NOTATION

~ Contact area ratio.

a Cross sectional area of cone

Bf Width of foundation

Bp Width of foundation or diameter of plate

b Dia. of asperities

Ce Compression index

Dm Constrained modulus (l/IIlv)

Dp Plate diameter

d Geophone spacing

On Normal deformation.

Onj Normal closure of a single discontinuity.

o nr Normal deformation of intact rock.

0nt Total normal deformation of intact rock and discontinuity.

Os Shear deformation.

E Young's modulus.

E+ Unload-reload modulus from pressuremeter tests

Ee Post-collapse modulus.

EdynamieDynamic modulus

Ee Re-Ioad modulus.

EI Intact modulus.

E· Initial modulus. 1

Eih Initial horizontal modulus

Eiv Initial vertical modulus

E· Joint modulus. J

ELM Lower bound modulus.

Em Mass modulus.

Eo Young's modulus at ground surface or foundation level

Es Secant modulus.

E Static modulus static

XXXVlll

E t Initial tangent modulus

EtSo Tangent modulus at 50% of the uniaxial compressive strength.

Ey Post-yield modulus.

EUM Upper bound modulus.

EO.001 Secant modulus at 0.001 % axial strain

EO.01 Secant modulus at 0.01% axial strain

€ Normal strain

€ c Cavity strain

€v Vertical strain

e Void ratio.

f Frequency

ff Fracture frequency

f1 Empirical factor (Stroud, 1988)

g Acceleration due to gravity

G Shear modulus

Gh Horizontal shear modulus

Gi Initial shear modulus

Go Shear modulus at the ground surface or foundation level

Gs Specific gravity

Gur shear modulus based on unload-reload cycles

Gv Vertical shear modulus

Gy Post-yield shear modulus

He Hammer energy

h Drop height

~ Plasticity index

IsSO Point load index

y Shear strain

J Rock mass factor (Hobbs, 1975)

~ Coefficient of earth pressure at rest

k Rate of change of Young's modulus E with depth

~ Normal stiffness.

~ Shear stiffness.

XXXIX

L Ratio EO.Ol : EO.OOI

1 Standard measure of cone displacement (lOOmm)

A Wavelength

M Mass of hammer

m Rate of increase in shear modulus G with depth

llly Coefficient of volume compressibility

N Parameter derived from the Standard Penetration Test.

Na Number of asperities per unit area

Nc Bearing capacity factor

N60 N corrected for 60% of the free fall hammer energy

n Porosity

p pressure

p'm Mean in-situ effective stress

Pf Creep pressure

PL Limit pressure

Po In-situ horizontal stress

Q NGI tunnelling quality index

v Poisson's ratio

P Bulk density

Pay Average settlement

Pcentre Settlement at centre of loaded area

Pd Dry density

P f Settlement of foundation

Po Surface settlement at a given radius from centre of plate

Pp Settlement of plate

Ps Settlement

Pw Density of water

P24 Plate settlement after 24 hours

R Creep ratio (Burland and Lord, 1970)

RMR Rock Mass Rating

Su Undrained strength

a c Uniaxial compressive strength.

xl

a t Brazilian tensile strength.

a y Yield stress

a 1 Major principal stress.

a 2 Intermediate principal stress.

a 3 Minor principal stress.

a t' Major principal effective stress.

a 2' Intermediate principal effective stress.

a 3' Minor principal effective stress.

e Phase difference

T Corrected phase difference

T Shear stress.

q Bearing pressure

qe Yield bearing pressure (based on the onset of yield)

'lnet Nett average bearing pressure

qy Yield bearing pressure (based on the establishment of Ey)

quIt Ultimate bearing pressure

q(ult)u Ultimate bearing pressure (undrained)

VI Relative volume of intact material

Vj Relative volume of joint material

V 0 Initial volume

V p Compressional wave velocity

Vr Rayleigh wave velocity

Vs Shear wave velocity

w Moisture content

WI Liquid limit

w sat Saturation moisture content

z Depth

xli

1.0 INTRODUCTION

The Chalk outcrops over approximately 15% of the area of England. Thus a

great many foundations will be constructed on this material each year.

With the exception of structureless chalk the strength of the rock mass is

sufficiently large for bearing capacity not to be a problem in most foundation

design. The major issue in the design of shallow foundations on chalk is

settlement since this controls the allowable bearing pressure.

The settlement of shallow foundations on chalk is controlled to a large extent

by the fractures which break the rock mass up into discrete blocks. Over most

of the chalk outcrop the dominant fractures occur in more or less parallel

sets that are generally sub-vertical and sub-horizontal. Weathering processes

result in these fractures being closely spaced near the ground surface and

becoming more widely spaced with depth. The compressibility of the rock

mass therefore likely to reduce with depth.

Pioneering work on the mass compressibility of the Chalk was reported by

Ward et al. (1968). This work related, however, to one stratigraphic,

palaeogeographic and tectonic setting. This is now seen to be atypical and, as

acknowledged by the original authors, the results are not applicable to the

Chalk outcrop as a whole. The Chalk is now known to be a highly variable

material (Clayton, 1983), with intact strength and stiffness properties which

may vary from those of a hard soil to a moderately strong rock.

The mass compressibility of the chalk has been studied extensively through

the use of plate loading tests. The typical load-settlement curve produced by

plate loading tests in chalk with essentially horizontal and vertical

discontinuities is distinctly convex in shape. This characteristic shape led

Burland and Lord to idealise the load-settlement relationship as essentially a

bi-linear curve such that the mass compressibility may be modelled using five

parameters. Burland and Lord suggested that the change in gradient of the

load-settlement curve is associated with yielding, but the nature of such

yielding was not postulated. The fact that some form of yielding occurs can

be seen in the results of unload-reload cycles in which it has been found that

deformation is essentially recoverable for pressures less than the yield point.

For pressures above the yield point significant non-recoverable deformations

are observed and the post yield modulus Ey is generally an order of

magnitude less than the initial modulus E j •

The magnitude of the yield stress is clearly important in foundation design

because of the significant increase in compressibility that is associated with it.

Records of full-scale foundations that have caused the chalk to yield are

limited to that of the sugar silos at Bury St Edmunds, Suffolk (Kee, 1974).

In total there are only six well documented records of the behaviour of

shallow foundations, of which only four deal with structured chalk. The

published record of plate loading tests are a little more extensive. The

majority of these are not reliable since the plate diameter used was too small

to provide representative results. As a result of the lack of reliable well

documented case records the mechanisms which control the mass

compressibility behaviour of chalk are not fully understood. The following r

aspects of mass compressibility behaviour are of particular interest in the

design of shallow foundations:

• The influence of intact mechanical properties.

• The mechanisms that give rise to yielding.

• The mechanisms controlling time-dependent settlement.

To date there has been no systematic investigation of any of these.

This thesis attempts to improve the fundamental understanding of some of

the factors controlling the mass compressibility of chalk and evaluate some of

the methods used in site investigation to predict foundation settlements. It is

impossible to gain a complete understanding of mass compressibility from a

2

single study such as this. The scope of this thesis is therefore limited to

investigating the influence of intact mechanical properties on the mass

compressibility of weathered structured chalk characterised by sub-horizontal

and sub-vertical discontinuities.

A total of nine large diameter (1.8m) plate loading tests of the maintained

load type were carried out at three different sites. A plate diameter of 1.8m

was chosen since this is close to the size of a full-scale foundation. Three

tests were carried out at each site. The sites were selected on the basis of the

following criteria:

• Each site must display different intact strength and stiffness properties

• At each site the intact mechanical properties must be reasonably

uniform with depth

• Each site must display similar dominant discontinuity orientations

(sub-horizontal and sub-vertical) and spacing.

The maximum load applied to the plate in each test was sufficiently large to

ensure that the post-yield load-settlement behaviour could be studied. Each

loading increment for a given test was maintained for 24 hours to ensure that

the rate of creep settlement had reduced to a reasonable level before

changing the load. Plate settlements were monitored throughout this period

to permit the study of short-term time-dependent settlement.

The rock mass at each test site was described in detail using trial pits (below

plate locations) and exposed chalk faces (adjacent to plate locations). The

intact mechanical properties of the chalk at each test site were measured in a

suite of laboratory tests.

At each test site surface-wave seismic tests and standard penetration tests

were carried out together with the classification of the rock mass using the

Mundford grading scheme (Ward et al. 1968). The results of these tests and

3

classification were used to predict the pre-yield settlement of the plate in an

attempt to evaluate their accuracy.

This thesis is organised as follows:-

Chapter 2 presents a review of the literature. This chapter is split into three

parts. The first part considers the deposition, diagenesis, geological structure

and weathering of the chalk. The second part considers the mass

compressibility of chalk and the third part the methods commonly employed

in site investigations to provide stiffness parameters for foundation design.

Chapter 3 describes the location of the sites selected for plate loading tests in

terms of geographical location, local topography and geology. In addition to

the location of test sites this chapter also describes the characteristics of the

rock mass at each site.

Chapter 4 describes the in-situ tests carried out at the test sites and the

laboratory tests on intact chalk. In this chapter the emphasis is placed upon

the plate loading test since this provides the best similitude to a real

foundation. The results of both the field and laboratory tests are given in this

chapter.

Chapter 5 presents a discussion of the work and results described in the

preceding chapters.

Chapter 6 presents the conclusions and recommendations for further work.

4

2.0 LITERATURE REVIEW

This thesis is concerned with the mass compressibility of fractured chalk, and

its influence on the settlement of shallow foundations. The compressibility of

any rock mass will be affected by the mechanical properties of the intact rock

and the geometry and mechanical properties of the discontinuities. These

factors are in tum affected by mineralogical composition, depositional

environment, diagenesis, tectonism and weathering. Any study of rock mass

behaviour therefore, should address these factors.

This survey of past literature is divided into three parts. The first part

considers the factors influencing the mass compressibility of chalk. The

second part considers the behaviour of foundations on the chalk and the third

part is a review of the methods commonly used in site investigation to

provide parameters for settlement predictions.

In the first part of the literature survey the following subjects relating to the

deposition, diagenesis, geological structure and weathering of the chalk are

reviewed:-

(i)

(ii)

The deposition of the chalk

This focuses on the chemical composition and the depositional

environment of the chalk in order to study the variability of the

rock material in terms of its physical properties.

Mechanical properties of intact chalk

Mechanical properties such as strength and stiffness are

examined since these are most likely to influence the

compressibility behaviour of the rock mass.

5

(iii) Discontinuities in chalk

The mechanisms for the formation of discontinuities in rock are

reviewed and these are used to examine the dominant

discontinuity patterns that exist in the chalk.

(iv) Weathering of chalk

The dominant weathering mechanisms which affect the chalk

are identified and the extent to which the rock mass is affected

these processes is examined.

The second part of this survey considers the mass compressibility of chalk.

The principal topics reviewed include:

(i) The influence of discontinuities on the compressibility of rock

masses is reviewed

The principal factors affecting rock mass compressibility are

identified. These are reviewed in relation to observed load­

deformation behaviour in the laboratory and in the field.

(ii) The load-deformation behaviour of fractured chalk

The characteristic load-deformation behaviour of chalk

dominated by sub-horizontal and sub-vertical discontinuities is

reviewed.

(iii) The behaviour of foundations

Published case histories of the load-settlement behaviour of

large-scale loading tests and full-scale foundations on fractured

chalk are reviewed.

The third part of this survey reviews the methods commonly employed in site

investigations to provide stiffness parameters for foundation design.

6

Since the discontinuities play such an important role in rock mass

compressibility it is normally impossible to predict settlements of foundations

using parameters provided by laboratory tests on intact chalk. Hence the

stiffness of the rock mass must be evaluated from in-situ tests and

observations. The methods considered include:-

• Pressuremeter tests

• Plate loading tests

• Geophysics

• Standard penetration test

• Visual assessment

By considering the above topics the current state-of-the-art for the mass

compressibility of fractured chalk may be reviewed.

2.1 Deposition, Diagenesis, Geological Structure and Weathering

To most engineers chalk is a white, weak and often friable material. It is

commonly associated with the south of England even though the outcrop

covers about 15% of the surface of England and extends into East Anglia,

Lincolnshire and the East Riding of Yorkshire. There are isolated outcrops in

Scotland and Northern Ireland but the material is different in character to

that found in England. For many years it was a commonly held belief

amongst engineers that the chalk is uniform material. Although this is

generally true with regard to the chemical composition of the chalk it is not

true for porosity. The large range of porosity (9 to 52%) stems from a wide

variety of depositional, diagenetic, and tectonic processes. This also gives rise

to a wide range of mechanical properties for intact chalk (Clayton

1983,1990a, Mortimore and Fielding 1990).

The chalk found near the surface is often highly fractured. Some of these

fractures may be considered to be 'primary' in that they result from

tectonism. These fractures will be found at depth as well as near the ground

7

surface and their geometry will be controlled by both regional and local stress

fields associated with the Alpine orogeny.

Many of the fractures found in the chalk near the ground surface may be

considered as 'secondary' since they result largely from weathering processes

such as stress relief and frost action. These fractures will reduce in number

with depth.

The near surface fractures in chalk are often modified as a result of

dissolution. This usually gives rise to rounded discontinuity walls and large

apertures. In extreme cases dissolution may result in larger features such as

pipe, dolines and sink holes. Such features may present a serious hazard to

foundation stability.

The combined effects of deposition, diagenesis, tectonism and weathering

give rise to a highly non-uniform rock mass. In order to characterise the chalk

it is necessary to gain a appreciation of how these factors influence the rock

material and the discontinuities.

Deposition and Diagenesis

The lowermost divisions of the formation (The Cenomanian - Table 2.1/1)

show the greatest variability in composition. The Cenomanian chalks are

typically greyish white in colour and characterised by rhythmically alternating

chalky limestones and interbedded marls in sequences similar to that found in

the Lias (Lower Jurrasic). This results from substantial quantities of clay

minerals and other clastic materials being introduced into the depositional

basin. However as the sea level rose in early Upper Cretaceous times the

amount of terrestrial detritus reduced such that the Turonian and Senonian

(Table 2.1/1) chalks are dominated by pelagic material, the bulk of which is

the debris from coccolithophorid planktonic algae (Haptophyceae), occurring

mostly in separate micron-sized plates or in their original rings known as

coccoliths (Hancock 1975). The calcium carbonate content of these chalks

8

Table 2.1/1 Zonal division of the Upper Cretaceous.

Stage Zone

Maastrichtian Belemnitella lanceolata

Belemnitella mucronata Campanian

Actinocamax quadratus Upper

Senonian Chalk Marsupites testudinarius

Santonian Micraster coranguinum

-------------Coniacian Micraster cortestudinarium

------------- -------------Holaster planus

Turonian Middle Terebratulina lata

Chalk Rhynchonella cuvieri

Cenomanian Lower Holaster subglobosus

Chalk Schloenbachia varians

9

(excluding marl bands and flints) generally exceeds 98% and rarely falls

below 96%. In terms of composition they display a high degree of uniformity.

Although most of the chalk has a more or less uniform composition it

exhibits a wide range of porosity. Clayton (1983) shows the range of dry

densities for English chalks to be 1.29 to 2.46 Mg/m3 which represents a

porosity range of 9 to 52%. The variation of porosity with biozone is shown

in Fig. 2.1/1. Fossil evidence suggests that the during the deposition of the

chalk substrate conditions must have been generally rather soft and therefore

inhospitable to many benthic taxa which are conspicuous by there absence.

Certain bivalves (eg species of Inoceramus, Pycnodonte and Spondylus)

became specially adapted to soft bottom conditions. The pelagic material

forming the sea bed would have behaved as a granular material in a loose

state which is characterised by high porosities. Indeed Scholle (1977) has

pointed out that carbonate mud may be deposited with as much as 70 to 80%

porosity. The major factor controlling the porosity of this material as it was

deposited would probably have been the grading, particle shape and degree

of bioturbation. Approximately half this pore space may have been lost by

dewatering during the initial stages of burial. Later diagenetic processes

during consolidation and cementation can reduce the porosity to less than

5 %. The fact that the chalk may retain very high porosities suggests that it

was subject to penecontemporaneous lithification. Lithification of carbonate

muds is often initiated as cementation at points of contact rather than

consolidation. Bathurst (1975) suggests that carbonate muds ·suffer very little

post burial compression because of this early cementation. Hence the rigidity

of weak carbonate rocks such as chalk may be attributed to mechanical

interlocking of grains, and the early cementing of the mass. It will be seen

from Fig. 2.1/1 that the range of maximum porosity in the white chalk (i.e.

the Turonian and Senonian Table 2.1/1) is similar to the range of porosity

within anyone zone. This tends to suggest that consolidation was not

significant in reducing the porosity of the majority of chalks. A

10

comprehensive discussion on the influence of consolidation on the

densification of chalk may be found in Clayton and Matthews (1987).

Hardgrounds are present in much of the chalk and are often sufficiently

extensive (e.g. Chalk Rock and Melbourn Rock) to provide geologists with

useful marker horizons for correlation purposes. These beds generally have

porosities of between 10 and 15% and are often nodular and show evidence

of intense bioturbation. Encrustation, phosphatisation and glauconitisation are

all features seen in these chalk hardground horizons. It is well known that

such hardgrounds were formed whilst the material was exposed on the sea

bottom during a break in deposition. During such time the carbonate mud

was exposed to intense biological reworking (Kennedy and Garrison 1975)

and significant cementation which brought about a significant reduction in

porosity. Hardground chalks have higher magnesium values than normal

white-chalk facies. This probably reflects early cementation by magnesium­

high calcite remobilised from shells of the richer benthos associated with the

more stable substrates.

Tectonism appears to be an important process in causing significant

reductions in porosity. The influence of tectonism on the hardness of chalk

has been noted by Strahan (1898), Mimran (1975), Clayton (1983) and

Clayton and Matthews (1985). The variation of dry density with the dip of the

chalk strata is shown in Fig. 2.1/2. It will be seen from Fig. 2.1/2 that there is

a increase in dry density with increasing dip. Since steeply dipping strata are

indicative of intense tectonic activity the data presented in Fig. 2.1/2 suggests

that tectonism brings about a reduction in porosity. Clayton and Matthews

(1985) demonstrate that as a result of tectonism large variations in porosity

can occur over relatively short distances (5 to 10km) within the same

stratigraphic horizon. The chalk in those regions affected significantly by

Alpine tectonism often exhibits evidence of compressional crushing of

microfossils whereas elsewhere in southern England such features are absent.

11

The chalk of Yorkshire and Lincolnshire is known to have a uniformly low

porosity; typically about 25% in contrast to the much more variable porosities

found in the chalks of southern England. Some chalks in Yorkshire contain

tests of microfossils that have been crushed and much of the Yorkshire chalk

contains blocky euhedral calcite crystals which appear to fill in excess of 50%

of the original void space (Bell et al 1990). Pressure solution and other late

stage solution features such as stylolites and thin beds (about 10mm thick) of

silty-clay material are common in the hard chalks found north of the Wash.

Hancock (1975) suggested that in the Santonian chalk (Table 2.1/1) of the

Yorkshire coast up to 50% of the original deposit may have been removed by

late-stage solution. Much of this material must have been available to

recement the deposit or to be redeposited as the blocky euhedral crystals

mentioned above. The mechanism for the reduction in porosity of the

Lincolnshire and Yorkshire chalks remains a subject of speculation. Scholle

and Kinsman (1973) and Scholle (1974) have proposed mechanisms based on

fresh water alteration and high geothermal gradients, but none as yet are

conclusive.

The above mentioned processes explain why the chalk, although generally

uniform in terms of composition, is extremely non-uniform in terms of

porosity. This makes the chalk rather unusual, since other sedimentary rocks

of similar lithology within the same formation generally show much greater

variations in composition. This feature of the chalk means that geotechnical

properties such as density, strength, stiffness and permeability of the rock

material may be related to porosity. The wide variation in porosity will give

rise to a similar variation in brittleness. The brittleness of the rock, together

with tectonism, will influence the pattern of discontinuities which cut the rock

mass. The properties of the rock material and the rock mass will be modified

by weathering agencies such as solution, stress relief and frost. Whereas the

geotechnical properties of the rock material may be simply related to porosity

the same cannot be said of the rock mass, although intact porosity no doubt

plays an important role.

12

Mechanical Properties of Intact Chalk

Chalk is typical of most weak rocks in that its intact properties can range

from soil-like behaviour to rock-like behaviour. The range of index properties

for the Chalk is shown in Table 2.1/2. Although no comprehensive systematic

study has been carried out on the mechanical properties of undisturbed intact

chalk there is sufficient published data to indicate that strength and

deformability are related to porosity. Porosity has been found to be an

important parameter in other rock types such as weak sandstones

(Dobereiner and de Freitas, 1983, 1986). Dobereiner and de Freitas (1983)

found that the most practical index parameter for assessing the strength and

deformability of weak sandstones was the saturation moisture content.

Duncan (1969) recognised the importance of this parameter for rock

classification purposes. For rocks of relatively uniform composition the

saturation moisture content may be related directly to porosity.

The unconfined compression and indirect tensile (Brazilian and point load)

tests provide a useful measure of rock strength which is relatively easy to

obtain, and can be used for comparative purposes. Indeed these represent the

most common of all rock mechanics tests (Bieniawski 1974). Mortimore and

Fielding (1990) found that there was a good correlation between intact dry

density and strength (Brazilian indirect tensile and uniaxial compression) for

white chalks. This suggests that a similarly good correlation should exist

between porosity and strength. Indeed porosity has been found to correlate

with both strength and elastic moduli for other rock types such as sandstone,

limestone and gypsum (Schiller, 1958, Kowalski, 1966, Morgenstern and

Phukan, 1966, Sirieys, 1966) and relationships have been proposed between

strength and porosity for other brittle materials such as ceramics (Knudson,

1959, Brown et al, 1964).

Published data on mechanical properties of intact chalk permit relationships

with porosity to be examined. The sources of these data are given in Table

2.1/2. The relationship between porosity and strength for the chalk is shown

13

Table 2.1/2. Range of Index Properties for the Chalk

Property Units Range

Dry density, P d Mg/m3 1.29 - 2.46 Porosity, n * % 9.00 - 52.00 Voids ratio, e * 0.10 - 1.10 Saturated moisture content, W sat* % 4.00 - 40.00

Calcium carbonate content % 55.00 - 99.00 Specific Gravity, Gs 2.69 - 2.71

Liquid limit, wL % 18.00 - 53.00 Plasticity index, ~ % 4.00 - 30.00 Liquidity index -2.25 - + 2.50

Point load index, ~(50) MPa 0.01 - 1.15 Uniaxial compressive strength, a c (Wet) MPa 0.30 - 22.00

(Dry) MPa 1.40 - 45.00 Brazilian tensile strength, at (Wet) MPa 0.14 - 2.20

(Dry) MPa 0.70 - 5.30

Initial tangent modulus, E t (Wet) GPa 4.60 - 13.90 Secant modulus, EtSO

(Wet) GPa 0.17 - 15.00 (Dry) GPa 0.17 - 18.00

* calculated from dry density assuming Gs = 2.70

(Data from Masson, 1973, Bell, 1977, Bonvallet, 1979, Clayton, 1973, Woodland et ai, 1988, Bell et al., 1990, Blight, 1990, Clayton and Saffari-Shooshtari, 1990, Kroniger, 1990, Mortimore and Fielding, 1990, Nienhuis and Price, 1990 and Varley, 1990)

14

in Figs. 2.1/3 and 2.1/4. It will be seen from Figs. 2.1/3 and 2.1/4 that

despite the scatter there is a distinct increase in strength with decreasing

porosity. The way in which the specimens were prepared and the test

conditions may contribute to this scatter. Most of the test specimens were cut

from block samples, but a number of different test specimen diameters

(38mm to 60mm) were employed for uniaxial compression tests. The ISRM

recommends (ISRM Commission, 1979) that a constant rate of loading be

employed for uniaxial compressive strength tests. Some of the results shown

in Fig. 2.1/3 are derived from constant rate of strain tests. However these

results do not stand out from the rest of the data suggesting that the

compressive strength of the chalk is not sensitive to the method of loading.

When specimens of chalk are tested dry (ie at natural moisture content or

saturated) they exhibit an increase in strength compared with their water

saturated counterparts. Fig. 2.1/5 shows the relationship between unconfined

compressive strengths derived from dry and wet specimens. It will be seen

from Fig. 2.1/5 that the strength is sensitive to moisture content. Generally

there is an increase in strength upon drying of between 40 and 50%, although

Masson (1973) reports ratios of dry:wet uniaxial compressive strengths

exceeding 4 for French chalks. It is well known that the strength of rock is

influenced by moisture content. Obert et al. (1946) showed small changes in

compressive strength for sandstone, marble, limestone and granite as the

moisture content varied, while a more comprehensive study by Colback and

Wiid (1965) defined relationships between uniaxial compressive strength and

moisture content for sandstone and shale. Colbeck and Wiid (1965) showed

that the strength of saturated rock was about half that of completely dry rock.

Similar relationships have been found by other authors (Krokosky and Hisak,

1968, Jumikis, 1966, and Ruiz, 1966). The reasons for variation of strength

with moisture content are still somewhat obscure. Hawkes and Mellor (1970)

suggest that the presence of water on internal surfaces of the rock produces

static fatigue, which may involve reduction of surface energy or fracture

energy (stress corrosion), bond modification, or atomic shielding. Little work

has been done on the influence of moisture content variations in chalk.

15

Mortimore and Fielding (1990) suggest that for dry densities of around 1.65-

1.70 Mg/m3 (porosity 39-37%) the strength is independent of moisture

content. The data presented in Fig. 2.1/5, however, indicate that low porosity

chalks still exhibit a significant gain in strength when dry. Some authors

favour testing only dry specimens (eg Mortimore and Fielding, 1990) because

of problems encountered during saturation. Indeed recent research on the

mechanical behaviour of cemented soils (Maccarini, 1988 and Bressani, 1990)

indicates that there is often significant damage to the soil structure through

isotropic stress oscillation when specimens are saturated under vacuum.

Clearly this could be significant when attempting to saturated high porosity

chalks since these will be only weakly cemented.

Most of the strength data shown in Figs. 2.1/3 and 2.1/4 were derived from

specimens cored perpendicular to the bedding. There is some evidence of

strength anisotropy in the chalk (eg Bell et al. 1990) but it is insufficient to

examine any relationship with porosity.

The relationship between Young's modulus and porosity for chalk is shown in

Fig. 2.1/6. Since the axial Young's modulus of intact rock varies throughout

the loading history it cannot be regarded as a constant for the material. It r

may be calculated in a number of ways. One of the most common is the slope

of the axial stress-axial strain curve at some fixed percentage, generally 50%

of the uniaxial compressive strength (EtSo)' The gradient of the initial portion

of the stress-strain curve is also used to determine a value of Young's

modulus (Et). Values of both EtSO and Et are shown in Fig. 2.1/6. Generally

Et

tends to be slightly (generally < 10%) less than EtSo, indicating that the

initial part of the stress-strain curve is concave in shape. At first sight the

concave shape might be attributed to bedding. However in many cases where

this has been observed the strains have been measured locally using electrical

resistance strain gauges. Bell et al. (1990) suggest that this behaviour is

associated with the closure of microcracks. This type of stress-strain

behaviour is typical of weak high porosity sedimentary rocks (Farmer, 1968)

16

Typical stress strain curves for chalk tested in triaxial compression are shown

in Fig. 2.1/8a. The strains were measured using electrolevellocal strain

gauges which are described by Burland and Symes (1983). It will be seen

from Fig. 2.1/8a that the stress-strain curves do not show the initial concave

portion though to be typical of weak porous rocks. This observation is

confirmed by plotting the secant modulus against axial strain as shown in Fig

2.1/8b. Fig. 2.1/8b shows the gradual drop in secant modulus with increasing

axial strain so that the final values are within 70-90% of the corresponding

initial stiffness. This behaviour may be a result of the method of saturation or

the method of testing.

The compression of chalk under conditions of uniaxial strain (zero lateral

strain) exhibits yielding which is typical of soils and weak rocks exhibiting a

bonded structure (Leroureil and Vaughan, 1990). Fig. 2.1/9 shows the results

of some one-dimensional compression tests on chalk of different porosities

reported by Leddra et al. (1993). It will be seen from Fig. 2.1/9 that chalks of

high porosity (ie n>35%) exhibits a relatively high stiffness up to the yield

stress beyond which there is a marked change in gradient of the void ratio­

stress curve indicating a significant reduction in stiffness. At stresses below

the yield stress the strains are generally recoverable (Leddra et ale 1993)

indicating elastic behaviour.

The yield stress marks the onset of destructuring (breakdown of bonded

structure). As the stress is increased sufficiently beyond the yield stress the

chalk becomes completely de structured and will behave like a granular soil.

This is seen clearly in Fig. 2.1/9 for the high porosity chalks. The initial steep

part of the void ratio-stress line represents the destructuring process. At high

stresses the completely destructured chalk shares a common void ratio-stress

curve which is independent of the initial porosity of the structured material.

This curve shows the stiffness of the material is increasing with increasing

stress which is typical of an uncemented soil.

17

The curve for chalk with 28% porosity in Fig. 2.1/9 shows no change in

gradient over the range of mean effective stresses used in these experiments.

However it will be seen from Fig. 2.1/9 that at the void ratio equivalent to

the initial porosity of this chalk the stiffness of the destructured material is

similar to that of the structured chalk and hence the yield stress will not be

marked by a noticeable change in gradient.

It will be seen from Fig. 2.1/9 that the yield stress increases as the porosity

reduces. This seems entirely reasonable since a high porosity is generally

associated with a weakly cemented chalk. The abruptness of yield appears to

be related to porosity in a similar manner.

Discontinuities in Chalk

In the mass the chalk is distinctly non-uniform, due principally to the wide

variety of discontinuity patterns displayed within this formation. Indeed the

fracture patterns often make the rock mass anisotropic (Nunn et al 1983,

Toynton 1983) and the degree and type of anisotropy varies significantly

across the extent of the outcrop. In geological terminology a discontinuity is

taken to include faults and joints. The separation between faults and joints

has frequently been associated with amount of displacement across the

fracture surface, and accordingly has been linked to scale. Joints are defined

as fractures formed to release stress within rock and along which there has

occurred insignificant displacement parallel to the fracture surface. Faults are

fractures along which there has been significant displacement parallel to the

fracture surface. Geologically they are considered to be different.

Geotechnically they share common characteristics such as low or zero tensile

strength and reduced shear strength compared with that of the intact rock.

Joints are relatively closely spaced structures. In small outcrops « 10m2)

there may be many joints belonging to the same set and often more than one

set may be present. On such a scale faults are generally rare, although some

faults may be closely spaced. In the chalk faults are generally rare except in

18

areas affected by intense tectonic activity. The major fold axes affecting the

Chalk of England are shown in Fig. 2.1/10. The Chalk has been affected by

only one major tectonic event; the Alpine orogeny. As the orogenic centre

was in continental Europe the intensity of effects in Britain decreases

northwards. Consequently, the only areas affected by complex structural

movements are south Dorset, the Isle of Wight and the Hog's Back in Surrey.

In these areas the chalk is found to be steeply dipping and cut by numerous

faults. Elsewhere the Chalk is affected only by relatively simple 'open' folds

such as the London Basin and the Wealden Dome. In the Chalk of East

Anglia, Lincolnshire and the Yorkshire Wolds, major folds are largely absent

and the chalk dips eastward at less than 10°. In general therefore the

principal discontinuities are joints.

Joint frequency is commonly a function of lithology and of bed thickness

(Harris et al, 1960; Hodgson, 1961, Price and Ladeira, 1981). Joint frequency

often increases markedly as the ground surface is approached. The chalk is

no exception, exhibiting closely spaced joints, often parallel to the bedding,

near the ground surface. Fault frequency is usually unrelated to lithology and

bed thickness. Joints are small two-dimensional structures. Minor joints are

generally restricted to single beds, particularly when the beds are folded or

the lithologies variable. Major joints usually cut several beds continuously. In

the chalk changes in depositional environment give rise to beds of different

porosity and hence brittleness. Such variations will often cause some joints to

be inpersistant and terminate at bedding discontinuities.

Slickensides are commonly associated with faults, but not with joints. In

tectonically hardened chalks it is common to find highly inpersistant

slickensided fractures associated with stylolites. In many rock types, joint

surfaces are irregular so that adjacent walls are completely interlocking. The

soluble nature of chalk however often results in the removal of small scale

irregularities which gives rise to limited point or ridge contact between

adjacent walls. The joints that appear most prone to this form of solution

19

weathering are generally those parallel with bedding, where the dip is sub­

horizontal.

In competent rock the type and orientation of any joint is governed by the

relative magnitudes of the effective principal stresses during propagation

(Hoek, 1968). The stress controls for joint development can be described

using Mohr's circles as shown in Fig. 2.1/11. A tension joint (extension joint)

will form under conditions of zero shear stress (a = 0°) where a 3' is tensile

and equal to the tensile strength of the rock (at) and where a l' is less than

3at• Where (a1 - ( 3) > 4at then a component of shear stress acts. For (a1 -

a 3) between 4a t and 8a t a hybrid joint will develop with a dihedral angle 2a

between 1° and 30° (Fig. 2.1/11). Joints forming directly due to shear (shear

joints) can only occur when a3' is compressive and then (a1 - ( 3) must be

greater than 8a t. Such shear joints will generally develop dihedral angles

greater than 60°. A more comprehensive explanation of these stress controls

may be found in Hancock (1985).

The patterns created by the development of joints within a rock mass range

from simple to complex. The range of possible patterns even for a single

tectonic event is immense and without establishing the development stages on

the basis of field data, prediction of jointing characteristics is likely to be in

error. Rawnsley et al (1990) suggest that there are three simple joint systems

that may result from a single event. These include:

(i) Polygonal systems

This system is developed where a2 = a3 = -at. Each joint forms

parallel to a 1 but otherwise is unrestricted. Such joint systems

are typical of lava flows and desiccated sediments and are not

seen in the chalk.

20

(ii) Grid lock systems

The term grid lock was suggested by Hancock (1985) for the

development of two orthogonal sets of joints as shown in Fig.

2.1/12a. In this case a 2 and a 3 are both tensile and 0 > Ia 2 -

a 31 > at. The development of one joint locally releases the

value of tensile stress perpendicular to it within a stress release

field that is proportional to the length of the joint (Pollard and

Aydin, 1988). As a2 and a3 are close in value, such a reduction

may reverse their relative magnitude such that the next joint to

form is perpendicular to the first (Hancock, 1985, Simon et aI,

1988). This process continues until stresses capable of causing

joint formation are released and an orthogonal joint system is

produced. In most cases a 1 is the overburden stress leading to

vertical joint formation.

(iii) Joint spectral systems

Joint spectral systems (Hancock, 1985) result from a gradual

increase or decrease in a 1 relative to a 3 leading to a full range

of joints from tensile to shear. An example of this type of joint

system is shown in Fig. 2.1/12b. The development of this system

is complex with each joint propagating through a constantly

varying stress field resulting from the applied stress and the

stress relief fields of neighbouring existing and developing

joints. Curved joints (over a few metres) are common in such

systems since a single joint may alternate through stages of

tensile, hybrid or shear failure in response to the changing

stress field. Within mixed lithologies different spectra may be

specific to individual beds, presumably due to differing strength

and brittleness characteristics. Variation can also occur in a

single lithology due either to lateral stress variation or changes

in material properties.

21

A comprehensive systematic study of jointing in chalk has yet to be carried

out that permits the classification of joint systems using the above modeL

However observations of jointing in chalk made by several authors (Cawsey

1974, Nunn et aI, 1983, Toynton, 1983) suggests that joint spectral systems

may predominate. Price (1966) suggests that in simply folded rocks two sets

of tension joints and two sets of shear joints should occur in the pattern

shown in Fig. 2.1/13. The model developed by Price has been successfully

applied by Hancock (1969) to the pattern of jointing and faulting in the

Jurassic limestones of the Cotswold Hills and by Cawsey (1977) to the Chalk

of southern England.

In areas away from complex structural movements the discontinuity pattern is

generally characterised by a set of sub-horizontal bedding discontinuities and

several sets of steeply dipping joints. Cawsey 1974,1977 studied the fracture

patterns in the Chilterns and found up to six well-defined sets of steeply

dipping joints (Fig. 2.1/14) the orientations of which can be predicted from

the regional and local geological structures using Price's model. Figure 2.1/15

shows the distribution of dip for these joints. Cawsey (1974) found that

although there was a wide variation in dip angles more than 50% of the

population sampled had dips greater than 800• Similar patterns of steeply

dipping joints have been observed in East Anglia (Toynton, 1983) and

Lincolnshire (Nunn et al 1983). In most cases four sets can be identified.

In areas affected by complex structural movements a different pattern of

discontinuities are observed. In the Central Downs of the Isle of Wight where

the monoc1inal fold gives rise to steeply dipping bedding only 14% of the

fractures have dips exceeding 800 (Cawsey 1974,1977, Fookes and Horswill

1970). Only two sets of discontinuities predominate in this area and both

show evidence of shear displacement (Fig. 2.1/16). The conjugate (diamond)

fissure pattern, commonly associated with tectonic folding in brittle rocks is

not well developed on the monoclinal fold (Fookes and Denness, 1969). This

tends to support the theory that this structure was produced by gravity in

22

response to deep seated block displacement rather than horizontal

compression.

In terms of discontinuity pattern the chalk appears to display a relatively high

degree of uniformity within a given tectonic environment. The most common

pattern is one of sub-horizontal bedding discontinuities combined with steeply

dipping often sub-vertical joints. The number of joint sets together with the

average spacing and persistence of each set can however give rise to a highly

non-uniform rock mass. Although the pattern of jointing may be predicted

with reasonable accuracy from the regional and local geological structures,

spacing and persistence cannot. Joint frequency is commonly a function of

lithology and of bed thickness (Harris et aI, 1960, Hodgson, 1961 and Price

and Ladeira, 1981). The presence of a hard ground in the chalk will influence

the joint frequency but otherwise the major factors controlling joint frequency

are tectonism and weathering.

The roughness or surface topography is likely to have a significant influence

on the mechanical properties of discontinuities. Indeed roughness is

considered to be of fundamental importance in controlling the shearing

resistance of rock discontinuities (Patton, 1966, Barton, 1973). One of the r

principal factors controlling the stiffness of individual fractures is the

geometry or surface topography of the joint walls. The surface topography of

a fracture will depend to a large extent upon the mode of fracture

development and subsequent processes such as shear displacement and/or

solutioning. Most fractures are formed either by tension or by shear. Tension

fractures typically have a rough surface with some rounded portions whereas

shear fractures often display a regular step-like pattern (Chemyshev and

Dearman, 1991) and may be slikensided. The surface roughness of

discontinuities will vary according to scale. Typically roughness is

characterized by:

23

(i) Waviness

Large scale undulations (wavelengths between 0.5m and 10m) which if

interlocked and in contact, cause dilation during shear displacement

since they are too large to be sheared off.

(ii) Unevenness

Small scale roughness that tends to be damaged during shear

displacement unless discontinuity walls are of high strength and/or

stress levels are low permitting dilation on these features. The surface

roughness at this scale appears to be determined by the dominant

grain size of the rock if it is less than a millimetre to a few centimetres

in fine and coarse grained rocks respectively.

If a tension fracture undergoes shear deformation at some stage the small

scale asperities are likely to be sheared off, particularly in fine and medium

grained rocks, resulting in a reduction in unevenness. In soluble rocks such as

chalk, percolating water will tend to remove the unevenness by dissolution,

increase the aperture and reduce the degree of contact between the joint

walls.

Little work appears to have been done in classifying joint surface topography

for different rock types despite the fact that the morphology of rock fractures

is fundamental to the understanding or modelling hydromechanical

behaviour. Gentier and Riss (1990), Hakami and Barton (1990) and Pyrak­

Nolte et al (1990) have studied the small scale morphology of fracture

surfaces in three dimensions of some igneous rocks. Bandis et al (1981) and

Aydan and Kawamoto (1990) have studied small scale surface topographies

for a variety of rock types in two dimensions. These sections indicate that in

two dimensions the surface topography at this scale may be considered as a

waveform in which symmetry, amplitude and wavelength are likely to be key

factors in classification. This "is essentially the basis o~spectral analysis

(Brown and Scholz, 1985). However for irregular and rough phenomena

showing partial correlations at a wide range of scales fractal analysis

24

(introduced by Mandelbrot (1977» is recommended (Gentier and Riss, 1990).

The bedding planes in limestone generally have a relatively rounded

topography compared with most of the others shown. The morphology of

these bedding plane asperities is associated with the depositional

environment. However it is likely that they have been modified by

solutioning. Over most of the chalk outcrop the bedding is sub-horizontal

and hence bedding discontinuities contribute much to the compressibility of

the rock mass beneath foundations.

For most of the chalk outcrop the dominant discontinuity system comprises

sub-horizontal bedding discontinuities and sub-vertical joints. The spacing of

these discontinuities is likely to be highly variable but may be controlled to

some extent by weathering processes and variations in lithology. In areas

where the chalk is dipping at angles greater than 10° the discontinuity system

will be more complicated and include steeply dipping joints together with

minor faults.

Most discontinuities in the chalk will be characterised by relatively rounded

asperities. The degree of contact between adjacent walls will be controlled by

largely by weathering processes such as stress relief and solutioning. These

processes are discussed in below.

Weathering

The principal weathering mechanisms which affect the chalk include:

(i) Stress relief

Stress relief results in the formation of new fractures and the

opening (increase in aperture) of existing fractures of tectonic

or depositional origin. The effects of stress relief die out with

depth or distance away from a cliff or valley side.

25

(ii) Frost or freeze/thaw action.

This results the mechanical disintegration of the chalk through

the formation of new fractures, opening of existing fractures

(regardless of origin) and the breakdown of the rock material.

In extreme cases the structure of the rock mass is completely

destroyed. Although the whole of the chalk outcrop has been

subject to type of weathering during the Pleistocene, some areas

are more seriously effected than others. In general those areas

such as southern England which were exposed to periglacial

activity throughout the Pleistocene tend to show the worst

effects which may extend 10's of metres below ground level.

(iii) Dissolution

The material adjacent to discontinuity walls undergoes

dissolution from percolating groundwater. This results in locally

increased apertures and the reduction in the degree of mating

of adjacent discontinuity walls. In extreme cases dissolution will

result in the formation of voids ranging in size from several

centimetres to several metres. The larger voids may collapse or

become infilled with soil. Such solution features are a common

problem in foundation design on chalk.

In the chalk these mechanisms combine to give rise to the characteristic

weathering profile shown in Fig. 2.1/17. Ward et al. (1968) observed the

weathering profile of the chalk in a number of large diameter boreholes at

Mundford, Norfolk and developed a engineering grade classification scheme

which is too a large extent controlled by the effects of weathering. This

classification scheme is shown in Table 2.3/9 and is discussed in detail in

section 2.2. It will be seen from Table 2.3/9 that zones A and B correspond

to Grade V (This also includes the additional Grade VI introduced by

Wakeling (1970), Zone C corresponds to Grades IV and ill and Zone D

corresponds to Grade IT (Grade I is defined on the basis of lithology and not

weathering).

26

The weathering zones may be identified on the basis of:

rock mass structure (ie presence of bedding and jointing);

discontinuity spacing;

and discontinuity aperture and infill

The zones shown in Fig. 2.1/17 do not have clearly defined boundaries. In

generally !~e boundaries between zones are gradational and the depth at

which a given zone grades into another is highly variable and is dependent

upon local conditions and recent geological history. The most noticeable

features of the weathering profile are the transition from unstructured to

structured chalk and the increase in discontinuity spacing with depth.

Examples of the variation of discontinuity spacing with depth are shown in

Fig. 2.1/18.

The chalk has been affected to a large extent by glacial and periglacial

conditions during the Pleistocene. The most significant is likely to be the

periglacial conditions since much of the chalk outcrop (ie southern England)

experienced these conditions more or less continuously during much of the

Pleistocene. Intact chalk is known to be susceptible to complete disintegration

by frost weathering (Tricart, 1956, Coutard et ale 1970, Lautridou, 1970,

Williams, 1980 and Jerwood et al., 1990). Near the surface the chalk is

intensely shattered and disturbed by frost heave such that the macro-structure

is destroyed. This structureless chalk comprising chalk fragments of all sizes

in a matrix of pasty matrix of remoulded chalk often resembles head, into

which it may pass upwards or laterally. However distinctive horizons such as

flint bands often continue through this material, although distorted by frost

heave, and prove that it is not transported material. Higginbottom (1966)

notes that the effects of frost shattering are most pronounced beneath

superficial deposits in the glaciated parts of the country, where they may

reach depths in the order of 15m. More than 12m of shattered chalk have

been recorded resting on the chalk beneath Quaternary tills and other

deposits of the Holderness Plain (Foster and Milton, 1976). In the south of

27

England frost shattering has been observed to depths in excess of 6m

(Higginbottom, 1966).

Frost shattering can cause selective opening of favourably orientated

discontinuities. Typically this affects discontinuities striking roughly with the

ground surface contours, through relief of expansive forces by downhill

displacement of the frozen layer. The effects often resemble small scale

cambering and may indeed be a related phenomenon. An example of this

form of frost shattering was found in the foundations for the first five piers of

on the western approach to the Medway bridge on the M2 in Kent. Loose

and open jointed chalk was encountered in which the principal discontinuities

had an average spacing of 460mm and apertures up to 50mm and most were

approximately parallel to the hillside, indicating an overall displacement

towards the valley centre (Higginbottom and Fookes, 1970). It was necessary

to grout to a depth of 6m below foundation level before the foundation

concrete could be placed (Kerensky and Little, 1964, and discussion 1965).

Most chalk excavations show very closely jointed material to depths in the

order of 30m (Higginbottom and Fookes, 1970). Where shallow deposits

overlie the shattered chalk, frost heaving has often produced involutions or

chalkland polygons and stripes, but where there are no such deposits the

chalk, though shattered is often not particularly disturbed (Williams, 1980).

Frost weathering accounts for much of the features of zones A and B and the

upper part of zone C. Within zone C the effects of frost shattering diminish

with depth such that the effects of stress relief may become more dominant.

There is no clear picture concerning the weathering of chalk in periglacial

times and hence it is impossible at this stage to predict the depth of frost

weathering for a given site. It is likely that there is a relationship between

intact porosity and the degree of disintegration of the chalk that would give

rise to different weathering styles. It has been noted (Williams, 1987) that

weathering is most clearly developed on the hardest (low porosity) chalks and

28

is relatively inconspicuous on weak chalks. However no systematic study of

chalk weathering can be found in the literature.

The effects of stress relief may be seen in the weathering profile particularly

in locations where frost shattering does not extend to great depth, although in

many cases it may be difficult to distinguish between them.

Stress relief generally produces joint-like fractures in brittle rock that are

commonly sub-parallel to the local topography, very local in extent and

normally have very little or no filling material (Nichols, 1980). They

sometimes crosscut tectonic fractures or become curved at the intersection

and terminate. Nichols (1980) suggests that these fractures are related to

lithology and proximity to free faces. For example in mixed sedimentary

lithologies such as limestone and shale the fractures are often contained

within a single unit and do not cross unit boundaries. The fracture

orientations are usually parallel to valley walls and in valley bottoms fractures

are parallel to the bottom surface. Hence the term 'topographic jointing' is

often used to describe these fractures.

The generally accepted idea, first proposed by Gilbert (1904), is that

expansion of rock masses when their confining pressures are reduced by uplift

and erosion finds relief in the development of cracks. This pressure-release

by erosion of a superincumbent load is seen mainly in granite, but also in

massive sandstone (Bradley, 1963), bedded sandstone (Currey, 1968) and

limestone (Kiersch and Asce, 1964). Stress relief through glacial erosion and

ice meltdown has resulted in significant fracturing of rock (Lewis, 1954).

Sheeting or topographic jointing is most common in massive brittle poorly

jointed rocks (eg Granite) In rocks that exhibit creep behaviour, such as high

porosity chalk stress relief jointing may not be well developed due to the

ability of the rock to deform in response to gradual stress changes. It is likely

that the style of stress relief jointing will vary across the chalk outcrop in

response to differences in brittleness related to wide variation in porosity. In

rocks masses which exhibit tectonic jointing the expansion may be taken up

29

by existing discontinuities hence the formation of 'topographic jointing' may

be severely restricted.

Fookes and Denness (1967) suggest that bedding the dominant control of

stress relief fracturing in chalk. They observed that if bedding is parallel to

the free ground surface, vertical unloading by erosion has a marked influence

on the development of fractures parallel to bedding.

The weathering profile may be modified by dissolution through the widening

of discontinuity apertures and the formation of larger scale features such as

pipes, swallow holes and dolines. Although the chalk has not developed the

huge cavern systems of the Carboniferous Limestone new cave systems are

being discovered every year, one of the most extensive occurring beneath

Beachy Head (Reeve et al. 1980).

In chalk as with other limestones, solution weathering is the process of

removing carbonate ions in an open system, thus permitting continuous

dissolution to take place. The extent of the dissolution and the removal of

the chalk as bicarbonate is controlled by a number of factors, the most

significant are likely to be:

(i) availability of water

(ii) availability of carbon dioxide

(iii) temperature

(iv) degree of saturation

(v) porosity

In addition to the carbon dioxide incorporated from the atmosphere,

percolating rainwater also takes carbon dioxide from the soil air (F ookes and

Hawkins, 1988). This additional carbon dioxide can increase the amount of

CaC03

that can be dissolved to over 30Omg/1 (Groom & Ede, 1972). Clearly

the soil profile can be a major source of chemical aggressivity. Atkinson and

30

Smith (1976) pointed out that in limestone terrains where a soil profile exists

the dominant solution activity takes place near the ground surface.

At low temperatures more carbon dioxide can be dissolved in water; for

example twice as much can be held at O°C than at 30°C. This has led some

authors (eg Higginbottom and Fookes, 1970 and Higginbottom, 1971) to

attribute the development of solution features in chalk to low temperatures

that existed during much of the Pleistocene.

Summary

The following points arise from this study of the deposition, diagenesis,

structure and weathering of the chalk:

• Most white chalk (Middle and Upper Chalk) exhibit a relatively

uniform chemical composition. However the processes of deposition,

diagenesis and tectonism have given rise to a wide range of porosity

(9% to 52%).

• The mechanical properties of intact chalk such as strength and

stiffness appear to be related to porosity. The strength and stiffness of

intact chalk can be expected to vary over two orders of magnitude.

• The discontinuity pattern in the chalk is dominated by sub-horizontal

bedding discontinuities and at least two sets of sub-vertical joints,

except in areas affected by intense tectonic activity. The primary

discontinuities (ie depositional and tectonically induced fractures) are

reasonably uniform within a given tectonic environment.

31

• Weathering processes result in significant changes to the structure of

the rock mass. The principal weathering processes include:

stress relief;

frost action;

and dissolution.

• The general weathering profile for the chalk is characterised by:

structureless chalk;

and structured chalk with discontinuity spacing increasing with depth

and aperture and infill decreasing with depth.

• The boundaries between zones within the general weathering profile

are gradational and their lateral and vertical extent are controlled by

local conditions and recent geological history.

• The weathering profile is dominated by the effects of intense freezing

during the Pleistocene. This has resulted in the formation of

structureless chalk and frost shattered chalk (ie closely to very closely

jointed chalk) to varying depths.

• Dissolution has resulted in the widening of discontinuity apertures,

modification of discontinuity surface topography and the development

of larger scale features such as pipe, swallow holes, dolines and caves.

These latter features can be a hazard to foundation stability in chalk

areas.

• There has been no systematic study of chalk weathering to date that

permits the identification of weathering styles associated with the wide

range of mechanical properties of this material.

32

Fig. 2.1/1

Fig. 2.1/2

Porosity (%)

Biozone 60 50 40 o

Bmue Min. Average Max.

Aqu

u~per M test

Calk Mea

Met

HP

Middle T lata

Chalk RhC

Lower Hsub

Chalk Sv

1.0 1.5 2.0 2.5 3-

Dry Density (Mg/m )

Variability of the chalk in terms of dry density and porosity (from Clayton and Matthews, 1987).

0

2.5-10

~ ,

Me •• , , , 20 ,

-.... • ,~'I' -C) 2.0- ffl. :E -- . \ •• 30 >->- -1'-........ • ... .-... tn . -tn 40 0 C "-

CD 1.5 0

C a.

>-Trend observed 50 ...

C by Mimran (1975) 60

1.0 0 30 60 90

Dip (degrees)

Results of density tests on chalk from West Surrey, Isle of Wight and the Isle of Purbeck (after Clayton and Matthews, 1987).

33

-as D.. :E -b

U

Fig. 2.1/3

-as D.. :E -b'"

Fig. 2.1/4

Porosity (%)

50~SO~ __ ~ ____ 4~0 ____ 3JO __ ~~ __ ~10L-__ ~0

40

30

20

10

0 1

III Dry chalk .,

+ Saturated chalk ..

1.2 1.4 1.6 1.8 2 2.2 2.4 2.S

Dry density (Mg/m3)

Variation of uniaxial compressive strength of chalk with dry density and porosity (Data from Masson, 1973, Bell, 1977, Bonvallet, 1979, Woodland et al., 1988, Clayton and Safari-Shooshtari, 1990, Bell et al., 1990, Blight, 1990, Kroniger, 1990, Mortimore and Fielding, 1990, Nienhuis and Price, 1990, and Varley, 1990).

Porosity (%) so 50 40 30 20 10 o

S,-~----~----~----~----~----~--~

5

4

3

2

1

0

El Dry chalk Gl

+ Saturated chalk '" '" '" '" '" '" ,,'" '" IY

" '" +,~-:" ",. £I ,-f

., '" , ... £1/ .... ~ + -~' ~ ......

~ +;...--~ - ~::r+:j:

1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6

Dry density (Mg/m 3

)

Variation of Brazilian tensile strength of chalk with dry density and porosity (Data from Bell, 1977, Bell et aI., 1990, Kroniger, 1990, Mortimore and Fielding, 1990 and Nienhuis and Price, 1990).

34

-. CO

D.. :E ---~ -CO .c Co)

~ ... C

bU

Fig. 2.1/5

-co D.. e" ---0 ~ -~ "0 0 E 0 ... C) C ~ 0 >

Fig. 2.1/6

50

40 0;. Dry = 2 0cSat

30

20 Ell

10

0 0 4 8 12 16 20 24

a Saturated chalk (MPa) c

Relationship between the uniaxial compressive strength of dry and saturated specimens of chalk (Data from Masson, 1973, Bell, 1977, BODvallet, 1979, Bell et aI., 1990, Blight, 1990 and Kroniger, 1990).

Porosity (%)

60 50 40 30 20 10 30

25 • Ei sat

• Et50 sat 20

• Et50 dry 15 • •• • • • 10

P !t . , •

5 . ~,.. •• •

0 1.00 1.50 2.00 2.50

Dry density (Mg/m3

)

Variation in stiffness of chalk with dry density and porosity (Data from Bell, 1977, Bonvallet, 1979, Woodland et al., 1988, Bell et aI., 1990, Kroniger, 1990 and Nienhuis and Price, 1990).

35

-CU D.. ~ --C")

b

b~ -

Fig. 2.1/7

Fig. 2.1/8

2000 B~

1500 A

N C

1000

500

O'~~------L-----__ ~ ________ -L ________ ~

o 0.02 0.04 0.06 0.08

Axial strain (%) Typical stress-strain behaviour for intact chalk (after Jardine et aI. 1985).

8 ,--------,--------,-------~--------~

-CU 6 D.. B CJ -t/) ::J -::J " 4 A 0 E c ... A Saturated samples c cu C tested undrained (,) CD 2 en B Part dried,

tested drained

o L-______ ~~ ______ ~ ________ ~ ________ ~

o 0.02 0.04 0.06 0.08

Axial strain (%) (bl

Variation of secant Young's modulus with axial strain for chalk (after Jardine et at. 1985).

36

Fig. 2.1/9

Fig. 2.1/10

1.0

0.8 48% porosity

0 .-... CU O.S ~

'"C .- 38% porosity 0 >

0.4 ...... _------- ---28% porosity

0.2 0 20 40 SO

Mean effective stress (MPa) -

Yield observed in chalks of different porosities when subjected to uniaxial (~ compression (after Leddra et al. 1993).

- Anticline

- - -- Syncline

1Skm

Chalk outcrop

Map of southeastern England showing the main structures affecting the Chalk.

37

a 2

a 2

a 3

(a)

Shear stress e\O~e e~~

-2T -T

Fig. 2.1/11

4T ~i,.~v.(e .................................... . ~ ....

.. . :/··· · ······· · ··· ...

2a = 45· ...

T Normal stress

aT

(b)

(a) Block diagram showing relationships between effective principal stresses and an extension fracture (E) and conjugate shear fractures (S) developed in a mechanically isotropic brittle rock. Stipple indicates the quadrants within which hybrid fractures form. (b) Composite failure envelope and Mohr circles constructed for 2a = 0, 45 and 60°, (T = tensile strength and <p = angle of internal friction. (after Hancock, 1985).

38

Fig. 2.1/12

Fig. 2.1/13

N

w ;------.;:: a.::.-----.... E

S N

(a) Grid lock system

w~----I"'''' .. ..----rE

s (b) Joint spectrum system

Rose diagrams showing typical grid lock and joint spectrum discontinuity patterns in rock (after Rawnsley et ale 1990).

-- Tension joints

- - - Shear joints

Typical system of jointing developed on the liob of a fold (after Price, 1966).

39

Fig. 2.1/14

U) CD 2 5 -- Jx

North Chilterns ca en 4 g II ::ii 015 -.-o 1 Z

U)

o

! South Chilterns Jy == .......... 4 ~ S! 3 J2 Jx' J3

.!2 II -.-2 oS

• 0 1 oeo Z o

90

Strike

90

Strike

J1

Base of chalk

180

180

(a)

Top of chalk

10km , ,

(b)

(a) Frequency polygons for dominant joint sets plotted against their strike for the Chalk recognised in the Chilterns. (b) Map of strikes of fracture sets observed in the northern Chilterns. (after Cawsey, 1977).

40

-fI. --->a () r::::: CI) ::J C" CI) ... --u..

Fig. 2.1/15

60,-----------------------------------

40

20

o~~~~~~~--~~--~~~~~~~~~

30 60 90 60 30

Dip (degrees)

___ J2 -+- Jx ~ J3 _ Jy --e-- J1 -8-- J4

Frequency polygons showing the variation in dip of the major joint sets recognised in the Chiltems (after Cawsey, 1977).

41

slickensided fault plane

Thrust fault

lolnt

joint

Flint band in plane of bedding

Joint

__ ~hearzone

3m I

Fig. 2.1/16 Block diagram illustrating orientation and frequency of joints and faults in the chalk near Culver Cliff, Isle of WIght (after Fookes and Horswill, 1970).

42

,

fragments Increases with depth and grades "' .. _~=:IP-- Friable to rubbly chalk. into structured chalk d· ~=~:::;:::::~<:::35~~= Bedding and jointing present.

\ Fractures are closely spaced ~-~-=-~~p-c=~- and generally Infilled with

With depth: .~..:::::::::::::::::=O- soil and chalk fragments. ~-h~~~ -- -{--

• fracture spacing increases D I --'--=:::--.

• Aperture decreases Blocky to massive chalk • amount of Inflll decreases Fractures are widely spaced and

generally tight

~----E----

Fig. 2.1/17 Typical weathering profile for the Chalk.

43

CD ()

~ :::J tJ)

"C C :::J o ... en ~ o a; Jl ..c

No. of vertical fractures per metre o 20 40 60 80 100 O.-------r-------.-------~------~------o .8 _.

- - - -v v

o Asham quarrey, Lewes

o Peacehaven cliff

• Taring Neville quarry, Newhaven

V Roedean cliff a CD 6 C

Fig. 2.1/18

Fig. 2.1/19

Variation of vertical joint spacing with depth for four sites in East Sussex (after Williams, 1987).

..-.. C 0 E ----CD > CD ...I

o

No. of horizontal fracture~ per metre

10 20 30 -25 ~--,.----r----

-30 v

A -35 AA

A

-40

Princess Quay

o BH1 v BH2 o BHa A BH9

Variation of sub-horizontal fracture spacing with depth for chalk, based on drillhole logs for Princess Quay, Hull (after Woodland et ale 1988).

44

2.2 Mass Compressibility

It is common practice in rock mechanics to emphasise the role of

discontinuities as the principal factor controlling the engineering behaviour of

rock masses. Since chalk is considered to be a weak rock, discontinuities are

thought to dominate the compressibility characteristics of the rock mass. The

description of the mechanical properties of rock discontinuities (particularly

rock joints), however, still appears to be a debatable subject. Although

several fundamental aspects of behaviour have been resolved qualitatively,

widely accepted quantitative procedures are limited. The literature reveals an

impressive number of articles documenting the shear strength of

discontinuities. The subject of rock joint stiffness and the influence this has

on the behaviour of the rock mass appears to be a relatively new subject and

hence has received less attention than shear strength.

Injluence of Discontinuities

Goodman (1976), Bandis et al. (1983) and Raven and Gale (1985) showed

that the introduction of a single discontinuity into a laboratory specimen of

rock perpendicular to the direction of the applied load will give rise to a

significant change in the stress-displacement characteristics when compared

with that of the intact rock. The deformation behaviour of a rock mass

involves the complex interaction of many factors. It is best understood by

separating the principal components of deformation. These include:-

(i) DEFORMATION BEHAVIOUR OF INTACT ROCK

Relative to the joints, intact rock is generally stiff. Its high

modulus is complemented by a low value of Poisson's ratio

which limits lateral expansion at moderate stress levels. It has

been shown in section 2.1 that the stiffness of intact chalk is

related to porosity.

45

(ii) NORMAL STRESS-CLOSURE BEHAVIOUR OF

DISCONTINUITIES

Goodman (1976) and Bandis et al. (1983) showed that

individually, joints display concave-shaped stress-closure curves

under normal loading. Fig. 2.2/1 shows that when the

deformation of the intact rock (0 nr) is subtracted from the stress

closure curve for the whole jointed block (0 nt)' a highly non­

linear, hysteretic stress-closure curve is obtained for the

individual joint (onj). Fig. 2.2/2 shows that cycles of loading and

unloading exhibited hysteresis and permanent set that

diminished rapidly with successive cycles. Fig. 2.2/1 shows that

non-mated discontinuities display greater closure at a given

normal stress that mated discontinuities. These characteristics

of normal stress-closure behaviour are thought to be typical of

most rock types.

(iii) DISCONTINUITY SHEAR BEHAVIOUR

Schieder (1976), Bandis et al. (1983) and Yoshinaka and

Yamabe (1986) showed that individually, joints display convex­

shaped stress-displacement curves under shear (Fig. 2.2/3),

usually accompanied by dilation which contrasts strongly with

the normal stress-closure behaviour. Fig. 2.2/3 shows that both

the peak shear strength and the displacement required to reach

the peak strength are scale dependent. The shear deformation

behaviour of discontinuities is strongly influenced by the

roughness of discontinuity surfaces and the degree to which

dilatancy is inhibited by the surrounding rock mass.

46

It will be seen from Figs. 2.2/1-4 that the stiffness of discontinuities in rock is

strongly stress dependent and influenced to a large degree by the mode of

deformation. The stiffness of individual discontinuities must therefore be

considered in terms of the normal and shear components of deformation.

Goodman et al. (1968) defined joint stiffness as the ratio between the

magnitude of the normal (a) or shear ('[) stress and the resulting normal (0 n)

or shear (os) deformation. This gives rise to two basic discontinuity stiffness

components:

Normal stiffness k n = ~ n

Shear stiffness ks = ; s

FACTORS AFFECTING NORMAL STIFFNESS

The principal factors which control the normal stiffness characteristics of

individual discontinuities include:-

(i) The area of contact between the two surfaces

(ii) The stress-strain characteristics of the intact rock

(particularly that adjacent to the discontinuity).

(iii) The presence of infill material and the mechanical properties of

this material.

(iv) The magnitude of the applied load.

47

Contact Area

When two rough surfaces are placed lightly together, the proportion of the

surface area in actual contact is almost zero. The entire normal load is

carried by a small number of point contacts. Under increasing normal load,

the point contacts enlarge by elastic deformation, crushing and tension

cracking, while deformation brings new regions into contact. Cook (1992)

suggests t~~t the normal stiffness of a discontinuity depends primarily upon

the contact area between the two surfaces.

The surface topography of joint walls will affect the distribution and

amplitude of asperities and the numbers of asperities in contact between the

two surfaces. This in tum will influence the asperity contact area, contact

stresses and the maximum joint closure. All these factors are dependent

upon the average stress level and whether the joint is mated or non-mated

(Fig. 2.2/4). Goodman (1976) made laboratory measurements of joint closure

as a function of normal stress on artificially induced tensile fractures in rock

cores. After the test with a non-mated joint, he observed that about 10% of

the area of the joint showed signs of crushing, representing the area of

contact between the two surfaces. Bandis et al (1983) have made an

extensive study of a variety of natural, unfilled joints with different degrees of

weathering and roughness in dolerite, limestone, siltstone and sandstone.

Impressions of contact areas between the two joint surfaces were made on a

polyester film 12 J..£m thick, inserted between the surfaces before loading.

These contact areas were found to be in the range of 40 - 70% of the total

joint surface area at the highest stresses. The effect of the film thickness

would be to overestimate the contact area, but natural joints are often mated

and hence would also be expected to have greater contact area than the non­

mated artificial joint measured by Goodman. Duncan and Hancock (1966)

made measurements of contact area at different applied stresses for a variety

of rock types including chalk, by inserting two sheets of "No carbon-Required"

pressure sensitive paper (manufactured by the National Cash Register

Company) between the two surfaces of the joints. The results of these

48

experiments are shown in Fig. 2.2/5. Despite the scatter of data points a

distinct trend of increasing contact area with applied stress may be seen. At

relatively low stress levels « 3MPa) the contact area was generally less than

10% of the total area. The contact area increases with stress level but

generally never exceeds 50% of the total area even at an applied stress of

20MPa. It is likely however, that Duncan and Hancock's experiments gave an

overestimate of contact area due to the effect of the thickness of the pressure

sensitive paper.

This trend of increasing contact area with normal stress was also observed by

Pyrak-Nolte et al. (1987) from measurements of joint closure, contact area

and void space geometry across natural joints in specimens of quartz

monzonite (50mm dia.) at room temperature and at 100°C. To study the

contact areas and void spaces between the surfaces of the joints, molten

Cerrosafe®, an alloy of lead, tin, bismith and cadmium that is totally molten

at 92°C, was injected into the joints at effective stresses of 3, 35, and 85 MPa

with a back pressure of 2MPa. The specimens were cooled to solidify the

alloy in place while the stress and back pressure were held constant.

Cerrosafe® has a surface tension of 400 roN /m, so that under a back pressure

of 2 MPa it will penetrate void spaces with apertures as small as O.4J,£m.

When the specimen was separated into two halves, precise metal casts of the

void spaces were found to adhere to one or the other of the two surfaces.

The void space and contact area could be mapped using optical techniques

and scanning electron microscopy. Fig. 2.2/6a shows the composite

micrographs of a portion of the natural joint at different effective stresses for

specimen E30. The black areas in Fig. 2.2/6a indicate the asperities of

contact. These contact areas appear to increase significantly with increasing

effective stress. Fig. 2.2/6b shows the relationship between joint closure and

normal stress and the relationship between contact area and effective stress

for the two specimens tested. At 3MPa the contact areas were 9 and 16% of

the total area. At 85MPa the contact area of the stiffest specimen (E32) had

asymptoted to a value of about 40% while that of the less stiff specimen

(E30) had reached a value of 30% and was continuing to increase. It is

49

apparent from Fig. 2.2/6b that even at stresses as high as 85MPa the joints

continue to close and significant void space remains within the joint. It will be

seen from Fig. 2.2/6b that the contact areas are of the same order of

magnitude as found by Duncan and Hancock (1966), Goodman (1976) and

Bandis et al. (1983).

It is clear from the above discussion that when a joint is subjected to

increasing stress that the asperities of contact increase in size and as a result

of joint closure new contact points are created. This behaviour is well

illustrated by a photoelastic investigation reported by Hyett and Hudson

(1990). In this study stress sensitive epoxy resins were cast against natural

joint surfaces in order to accurately reproduce their surface morphology. For

mated joints the number of new point contacts was greater for the rough than

for the smooth joint as the applied stress was increased. The pre-existing

point contacts for the smooth joint tended to spread laterally with increasing

applied stress. For mismatched joints the number of joint contacts created as

the load was increased was few, and to compensate for this the stress

transmitted across individual contacts was higher. As with the interlocking

cases individual contact zones spread more rapidly for the smooth joint than

for the rough joint. The smooth mismatched joint probably best models the

bedding plane discontinuities in chalk that have had material removed by

solution.

A number of mathematical models have been proposed to describe joint

closure under an applied loading which take account of joint surface

topography (Greenwood and Williamson, 1966, Swan, 1983, Brown and

Scholz, 1985,1986). The theories used to model joint compliance are based

on elasticity and yet they predict decreasing compliant behaviour with

increasing stress which is observed in experiments involving both elastic and

plastic deformations. These models are clearly deficient in that they fail to

recognise the fact that asperities of contact are likely to undergo yield or

collapse particularly in weak rocks None of these theories of joint closure

50

takes account of the interaction between asperities and the related

deformation of each elastic half space bounded by the joint surfaces.

Stress-Strain Behaviour of Discontinuity Wall Rock

The surface topography of the two joint surfaces and the degree to which the

two surfaces are mated will clearly control the initial contact area. As the

load is applied perpendicular to the joint, load is transmitted across the

asperities of contact and high contact stresses are generated at these points.

The stress-strain characteristics of the intact rock forming the joint wall will

control the response of the asperities of contact to these induced stresses. In

strong rocks of low porosity its is likely that the asperities of contact behave

elastically at engineering stress levels, and that in such a case the normal

closure of the joint will be recoverable upon unloading. Sun et al (1985)

carried out experiments on artificially produced discontinuities in rock types

of relatively high strength (Granite (Jc > 200 MFa and Slate (Jc > 300 MFa).

They found that for the first loading of an un-mated joint up to 72% of the

deformation was recoverable, whilst in repeated loading on other un-mated

joints this figure could be as high as 92%. The high degree of elastic

response of these materials, particularly at the low stress levels employed « r

0.1 * (J c), will have a significant influence on the observed behaviourial trends.

Brown and Scholz carried out joint closure studies on a £ 80 grit roughened

marble surface which had been loaded against a flat surface over four cycles

to 10MPa. The closure data show a large permanent set for the first cycle

and recoverable deformation for subsequent cycles. Scanning electron

microscope studies show definite plastic flow at the higher asperity tips. For

the flattened tips yielding occurred resulting in plastic flow until the contact

area became sufficient to support the load elastically for the remaining cycles.

This plastic behaviour was not observed as a dominant joint closure

mechanism in similar experiments on rocks such as quartzite and granite in

which the individual grains are stiffer than the calcite forming the marble.

51

In weak rocks even at relatively low applied stress levels it is likely that

asperities of contact could undergo yield and hence behave in a plastic

manner. In such a case the joint closure would not be recoverable upon

unloading. In highly porous weak rocks such as chalk the asperities of contact

may undergo pore structure collapse which may result in a sudden change in

gradient of the joint closure-normal stress curve. Pore structure collapse has

been observed in intact specimens of chalk (Leddra et al. 1990). Compression

tests on artificial discontinuities in chalk have demonstrated that in uniaxial

strain the discontinuity exhibits yielding at average normal stresses much less

than that of intact chalk (Matthews and Clayton, 1992). This behaviour

contrasts with the characteristic concave stress-closure curves observed by

Goodman (1976) and Bandis et al. (1983) for other rock types. Wake ling

(1975) and Matthews and Clayton (1992) showed that chalk exhibits this

behaviour for smooth fractures with contact area ratios (contact

area/specimen cross sectional area) greater than 90%. The yielding

behaviour was observed for specimens with contact area ratios less than 30%.

Clearly the normal stiffness of discontinuities depends upon an

interrelationship between the stress-strain behaviour of the intact rock

forming the asperities and contact area.

Weathering of discontinuity walls is likely to have a significant effect on the

stress-strain behaviour of asperities which in turn will influence the normal

stiffness of the discontinuities. Chemical decomposition of rock forming

discontinuity walls is likely to reduce the stiffness of asperities of contact.

Dissolution of discontinuity walls will reduce the contact area.

In the chalk the wide range of porosity gives rise to a wide range of intact

mechanical properties which may have a significant influence on the

compressibility characteristics of the discontinuities.

52

Discontinuity Infill

The presence of infill material within the discontinuities will effectively

increase the contact area between the two surfaces and reduce the void space

and hence is likely to have a significant effect on the stiffness characteristics

of the fracture. The influence infill has on the compressibility of a

discontinuity will depend to a large extent upon the degree to which the void

space is filled and the mechanical properties of the infill material. If a joint is

completely infilled the stiffness will be controlled by the thickness of the infill

relative to the amplitude and distribution of asperities together with the

mechanical properties of the infill material. If the infill thickness is greater

than the maximum amplitude of the asperities then there will be no contact

between the asperities and the stiffness will be controlled by the mechanical

properties of the infill until joint closure brings the asperities into contact. If

the thickness of infill is such that there is asperity contact then it is likely that

the asperity contact stresses will be reduced through load shedding onto the

infill since deformation of the infill may be inhibited by lack of drainage

(cohesive infill) or by the lack of sufficient void space if the infill is dilatant.

In such a case the stiffness of the joint may be increased and in the case of a

cohesive infill result in time-dependent joint closure. If the infill is loose and

free draining it is likely that the presence of the infill will have little effect on

the joint stiffness provided there is asperity contact. Of course in all these

cases ultimately joint closure will be inhibited by both the increase in asperity

contact area and the increase in stiffness of the infill as it consolidates. This

assumes the infill behaves as a soil. In some cases joint infill can be cemented

and hence may behave as a rock. In such cases it is likely that the joint

stiffness would be increased.

If the discontinuities are only partially infilled the normal closure will

proceed as if the infill were not present until the volume of the remaining

void space approaches that of the infill. In such a case there would be

interaction between asperity contact and the infill in the manner described

above.

53

Little work appears to have done on the influence of compressible infill

material on joint stiffness. Goodman (1970) tested artificial joints infilled with

crushed mica in direct shear. The shear stiffness of the joint decreased slowly

with degree of joint filling until the infill thickness was greater than 0.8 times

the amplitude of the asperities at which point the stiffness remained constant

at twice the stiffness of the infill material. This represents one of the first

investigations of the influence of infill on the mechanical properties of joints.

This and subsequent investigations (Lama 1978, Papaliangas et aI., 1990,

Pereira, 1990, Phien-wej et al., 1990 and Xu et al., 1990) have all focused on

shear strength. In terms of normal stiffness the principal factors relating to

the influence of infill would include:

(i) Stiffness of infill material

(ii) The degree of infilling (ie whether the joint is partially or

totally infilled) and the thickness of the infill.

(iii) The degree of rock to rock contact across the joint.

Most joints in chalk are either free of infill or only partially infilled with

chalk fragments. The rock to rock contacts across the joint are generally

dominant and hence the infill will have a negligible effect on the joint or

mass stiffness.

FACTORS AFFECTING SHEAR STIFFNESS

Shear stiffness is unlikely to make a significant contribution to the

compressibility of chalk which is dominated by sub-horizontal and sub-vertical

discontinuities. It is the normal stiffness of the horizontal and sub-horizontal

joints that will tend to control the deformation of the rock mass in response

to a vertical foundation loading. A comprehensive review of the factors

affecting shear stiffness is therefore beyond the scope of this thesis. For this

reason the factors affecting shear stiffness are only outlined below.

54

The principal factors affecting the shear stiffness of a single discontinuity

include:

(i)

(ii)

The roughness of the discontinuity walls

The stress-strain behaviour and shear strength of the

discontinuity wall rock

(iii) The presence of infill or materials coating the discontinuity

walls

(iv) The magnitude of the applied load

Within a rock mass the discontinuities may be so orientated with respect to

the applied stresses that mode of deformation is one of shear displacement.

In such a case the degree of interlocking of asperities, the amplitude of

asperities, the shear strength of asperities, the freedom for dilatancy and the

presence of infill will be important factors in controlling the shear stress­

displacement characteristics of the discontinuity. The shear stress­

displacement curves shown in Fig. 2.2/3 suggest that joint behaviour in

translational shear under constant normal stress may vary from 'brittle' to

almost 'plastic', depending on the size of the joint. Barton et al. (1981) in a

survey of peak shear stiffness found that for a given normal stress, shear

stiffness was inversely proportional to discontinuity length. Peak shear

strength, peak shear displacement and shear stiffness are all depicted as scale

dependent parameters (Barton and Bandis, 1982). In a rock mass, however,

translational movements along discontinuities are unlikely to occur under

conditions of constant normal stress. Since translation often involves dilation

the normal stiffness of the discontinuity will cause the normal stress to

increase. Rock joints tested under constant normal stiffness conditions still

produce convex shear stress-displacement curves.

55

FACTORS AFFECTING THE COMPRESSmILITY OF A FRACTURED

ROCK MASS

It is clear from the above discussion that the stiffness characteristics of

individual discontinuities are associated with interaction of several factors

such as contact area, infill and the mechanical properties of the intact rock.

The compressibility characteristics of a rock mass however involves the

interaction of multiple discontinuities and intact rock. The principal factors

controlling the compressibility of a rock mass subject to a foundation loading

must include:-

(i)

(ii)

The stiffness characteristics of the discontinuities present

within the zone of influence of the applied load.

The orientation of the discontinuities relative to the direction of

the applied load.

(iii) Confining pressure.

(iv) The number of discontinuity sets present.

(v) The spacing of discontinuities relative to the size of the loaded

area.

(vi) The stress-strain characteristics of the intact rock.

(vii) The magnitude of the applied load.

The stiffness characteristics of individual discontinuities will play a major role

in controlling the compressibility of a fractured rock mass. The factors which

influence the stiffness of discontinuities has been discussed earlier. The

contrasting shapes of the normal stress-closure and shear stress-displacement

56

curves will clearly give rise to different modes of rock mass deformation for

different orientations of discontinuities relative to the direction of applied

load.

Orientation of Discontinuities Relative to the Direction of Applied Load

Barton 1986 defined three characteristic modes of load-deformation

behaviour associated with the orientation of the discontinuities relative to the

direction of the applied load based on the dominant deformation mechanism.

These are summarised in Table 2.2/1 and Fig. 2.2/7. Fig. 2.2/7 shows the

three simple rock mass models considered by Barton (1986) subjected to a

vertical applied load.

Table 2.2/1. Characteristic load-deformation behaviour for rock masses.

(After Barton, 1986)

Case Dominant Shape Hysteresis Lateral mode of of load- expansIon Deformation Defm. curve

A Normal Concave Small Small

B Normal & shear Linear Moderate Moderate

C Shear Convex Large Large

In case A the dominant discontinuity orientations are horizontal and vertical.

In this case the normal closure of the horizontal fractures will be the

dominant deformation mechanism. Since, individually, joints display concave­

shaped stress-closure curves under normal loading (Fig. 2.2/1) it seems

reasonable to assume that the load-settlement curve for the rock mass

displays a similar concave shape (Fig 2.2/7a). Brown and Trollope (1970) and

Chappell (1979) demonstrated this concave behaviour in laboratory-scale

model tests with fractures parallel and perpendicular to the applied major

57

principal stress (see Fig. 2.2/8). This characteristic concave load-settlement

curve has been observed in large diameter (l.5m) plate loading tests on

Bunter Sandstone at Heysham, Lancashire (Hobbs, 1973) which is dominated

by sub-horizontal bedding discontinuities and steeply-dipping joints (Meigh et

al. 1973). Once the joints have closed sufficiently the stiffness of the intact

rock will begin to dominate the load settlement behaviour. The interaction

between joint closure and intact rock deformation depends upon the relative

stiffness of the joint and the intact rock together with the strength of the

asperities.

In case B (Fig. 2.2/7b) the orientation of the discontinuities are such that

the deformation of the rock mass must involve both normal closure of

fractures together with shear displacement. Individually joints display a

convex-shaped stress-displacement curves under shear, usually accompanied

by dilation (Fig. 2.2/4). Barton (1986) suggests that a linear load-settlement

relationship is applicable to this model (Fig. 2.2/7b) resulting from the

superposition of normal closure (concave stress-closure curve) and shear

displacement (convex stress-displacement curve) components of deformation.

Cramer et al. (1984) describe large-scale in-situ block tests performed on

columnar basalt. When the load was applied perpendicular to the basalt r

columns the load-deformation curve was more or less linear. The concave

behaviour associated with normal joint closure was absent during loading but

was evident during unloading suggesting that a combination of shearing and

normal joint closure was occurring. However it is unlikely that the magnitude

of normal closure will have been balanced by the magnitude of shear

displacement such that a linear load-settlement curve results. The shear

displacement is likely to be associated with dilatancy (depending on the

roughness of the joint walls). DilatanCY, however) may be inhibited by the

normal stiffness of the fractures on which the shear displacement is taking

place. This is likely to reduce the amount of shear displacement and hence

the expected load-settlement curve for the rock mass is more likely to be

concave than linear. Load deformation curves obtained from the triaxial tests

on hexagonal columns (with their long axes horizontal) reported by Brown

58

(1970) show linear stress-strain curves only at high confining pressures. When

unconfined the load deformation curves where distinctly concave. Since

deformation was measured externally in these tests this behaviour may be

associated with bedding at the platens. However since the concave curvature

dominates the stress-strain behaviour up to failure it seems unlikely that

bedding during the initial stages of the tests is masking much of the rock

mass behaviour.

In case C (Fig. 2.2/7c) the discontinuities are all inclined relative to the

direction of the applied load such that the dominant deformation mechanism

is one of shear displacement. Barton (1986) suggests that a convex load­

deformation curve will be associated with such a geometry since the shear

stress-displacement curves for individual joints are typically convex. Convex

load-deformation curves were observed by Brown and Trollope (1970)

Chappell (1979) and Yoshinaka and Yamabe (1986) in model tests with

blocks in this configuration. Normal closure will however, still play an

important role in the deformation of this rock mass since the dilatancy

generally associated with the shear displacement will be controlled by the

normal stiffness of the inclined joints which in tum influences the shear

stiffness of these fractures.

Confining Pressure

Laboratory-scale model tests have shown that the stress-strain behaviour of

fractured rock masses is dependent upon confining pressure (Brown, 1970,

Brown and Trollope (1970) and Yoshinaka and Yamabe, 1986). The

application of high isotropic confining pressures is likely to bring the blocks

in the model closer together thus increasing the degree of contact across

joints. This would clearly result in the stiffness of the rock mass increasing.

High confining pressures however, are not relevant to the load deformation

behaviour of a rock mass subjected to foundation loading near the surface.

59

Confining stress is likely to influence the shear stiffness of discontinuities

more than it does the normal stiffness since increasing confining stress will

inhibit dilatancy. It is unlikely that confining stress will have a significant

influence on the load-settlement behaviour of near surface weathered chalk,

particularly in cases where the sub-vertical discontinuities are open.

Number of Discontinuity Sets

Discontinuities normally occur in sub-parallel sets which are controlled by

tectonism, cooling and stress relief. The number of sets present within a rock

mass will influence the dominant mode of deformation and the average

block size and shape. The smaller the average block size the greater the mass

compressibility.

Discontinuity Spacing relative to the size oj the loaded area

In general the dimensions of a loaded area will control the volume of rock

which experiences an increase in stress. The geometry of the discontinuities

within the rock mass will have a strong influence on the distribution of

stresses associated with the load applied by a foundation which complicates

this model. However if the dimensions of the foundation are less than the

average discontinuity spacing the discontinuities will have little effect on the

load-settlement behaviour. This is rarely the case for weathered near surface

rocks. In most cases the discontinuity spacing is much less than the

dimensions of the foundation and hence the stiffness of these fractures will

dominate the load-settlement behaviour. The rock mass compressibility will

therefore be sensitive to the number of discontinuities that occur within the

zone of influence of the loaded area.

The influence of discontinuity spacing upon mass compressibility may be

explored mathematically by considering the rock mass as a composite

material made up of joints and intact rock. The modulus values for joints

and intact rock may be considered as lower and upper bounds respectively for

60

the resultant composite (the rock mass) when considering the criteria of

either equilibrium or compatibility (Voigt, 1928, Reuss, 1929). When

determining the upper and lower bound moduli for a composite material a

simplistic yet correct model can be employed. If the joint modulus Ej is much

smaller in magnitude than the intact modulus E and the joints are orientated

perpendicular to the direction of the applied load (Fig. 2.2/9a) then it is the

modulus of the joints which control the deformation of the rock mass and this

rise to a lower bound value of mass modulus. If however the joints are

parallel to the direction of the applied load (Fig.2.2/9b) it is the intact

modulus which tends to control the deformation rock mass and gives an

upper bound mass modulus. If the joint set is inclined to the direction of the

applied load (Fig. 2.2/9c) the mass modulus is a combination of the upper

and lower bound moduli and in the shear modulus of the joint set must be

considered. A comprehensive description of the shear modulus effect is given

by Chappell (1979).

The lower and upper bound moduli for a rock mass of the type shown in Fig.

2.5/16a may be defined by :

(i) Averaging the elastic moduli which is equivalent to assuming a

uniform strain on the boundary of each region, that is satisfying

compatibility (Voigt 1928).

where:

Em = Lower bound modulus

By = Intact modulus

Ej = joint modulus

VI = relative volume of intact material

V. = relative volume of joint material J

61

(ii) Averaging the compliancies which corresponds to the

assumption of uniform stress on the region boundaries (Reuss

1929).

where:

BuM = Upper bound modulus

The effect of discontinuity spacing on the mass modulus may be depicted

using equation (3) to determine a lower bound modulus of a varying number

of joints with different ratios of joint to intact material moduli. The case with

the joints open to 1.0mm, aligned perpendicular to the direction of the

applied load is shown in Fig. 2.2/9a. This represents the geometry often

found in chalk. It will be seen from Fig. 2.2/10 that as the modulus of the

joint approaches that of the intact material (ie E/EI = 0.5 in Fig. 2.2/10) the

number of joints has little effect on the ratio of mass modulus to intact

modulus. The implication is that for the softer type of materials such as soils

the joint system does not cause an appreciable difference between the mass

modulus and intact modulus. The most significant feature of this model

however is the extremely rapid drop in Ern/E! with the introduction of only a

few (1 to 8) fractures per metre when the ratio of E/E! is less than 0.005.

When the fracture frequency exceeds about 10 per metre the mass modulus

becomes relatively insensitive to increasing numbers of joints.

The relationships shown in Fig. 2.2/10 assume a constant joint aperture of

1mm. If the ratio E/EI is kept constant and the aperture allowed to vary the

set of curves shown in Fig. 2.2/11 results. The greater the aperture the

greater the mass compressibility. Wakeling (1975) carried out some

laboratory compression tests on cylinders of chalk stacked end on end to

simulate jointing. The results of these tests shown in Fig. 2.2/12 exhibit a

similar trend to that predicted by the model. Since the ends of each cylinder

were smooth the aperture was probably less than O.5mm. The values of rock

mass factor j are much lower than those shown in Fig. 2.2/11 for the same

62

joint frequency. This suggests that the ratio E/E. for this chalk was much less

than 0.005.

This model assumes that the aperture is of constant width for each joint. This

is rarely the case in practice. Bandis et al (1983) suggest that partial contact

between blocks would account for a low mass modulus of apparently rigid

blocks, since the stresses between the blocks would be transmitted by

asperities of contact which vary in size, shape and height. Pursuing this line of

reasoning Hobbs (1975) derived a simple ideal model in which uniform

contacts are regarded as miniature plate loading tests. The rock mass factor j

which is equivalent to Em/EJ is obtained from the expression given by:-

Where:-

j = Rock mass factor

ff = Fracture frequency

b = Dia. of asperities

Na = Number of asperities per unit area

v = Poisson's ratio

The contact area ratio ~ (ie contact area per unit area of joint wall) is given

by:-

1tb2N A. = a * 100%

J 4

Figure 2.2/13 shows the relationship between rock mass factor j and fracture

frequency for different contact area ratios ranging from 5 to 36.5%. It will be

seen that the curves are a similar shape to those derived from the above

model and that as the contact area ratio increases the curves become flatter

and less sensitive to fracture frequency. However, as one might expect from a

model based on contact, Hobb's model is sensitive to the relative magnitudes

of Na and b. For a given contact area ratio, Poisson's ratio and fracture

63

frequency, the Rock Mass Factor j increases with an increasing number of

asperities per unit area (see Fig. 2.2/14). This implies that as the asperity

contact area reduces the rock mass compressibility is reduced. Another

feature of Hobb's simple model is that it becomes ill-conditioned for contact

area ratios greater than 36.5% and yield rock mass factors greater than unity

implying that the fractures cause the fractured rock mass to be less

compressible than the intact rock.

The discontinuity pattern in chalk outside the areas affected by intense

tectonic activity is generally dominated by sub-vertical and sub-horizontal

fractures and hence the simplistic models discussed above may be applied to

the chalk mass. The rock mass factor j is plotted against joint frequency in

Fig. 2.2/15 for four Chalk sites. All are based on large diameter plate tests

with the exception of the Killingholm site where the pressuremeter was used.

The results exhibit a similar trend to that predicted by the models discussed

above indicating the importance of joint frequency, aperture and asperities in

controlling the mass compressibility. Ward et aI. (1968) recognised the

importance of joint frequency and aperture in the development of an

engineering grade classification (Table 2.3/5) for predicting the mass

compressibility of Chalk at Mundford. This classification system is discussed

later.

These simple models relating joint frequency to mass modulus assume that

the ratio E/EI remains constant for the system of joints under consideration.

However the tests on individual fractures (Goodman, 1976 and Bandis et aI.

1983) have shown that Ej is highly stress dependant. In reality the stiffness of

the joints beneath a loaded area is likely to change with depth as a result of

the distribution of applied stresses. This will clearly modify the relationships

discussed above. A load applied over a large area will influence a larger

volume of rock than the a load of similar magnitude applied over a smaller

area. In the latter case only few discontinuities may be present within the

zone of influence of the applied loading whereas in the former case there is

scope for many more fractures to be present which clearly would give rise to

64

different mass compressibility characteristics. Therefore it is not simply the

fracture spacing that affects the mass compressibility but the spacing in

relation to the size of the loaded area. The situation is further complicated by

the fact that the orientation, persistence and aperture of the joints will have a

strong influence on the distribution of stresses within the rock mass

associated with an applied load.

Stress-strain Characteristics of the Intact Rock

The stress-strain characteristics of the intact rock can be expected to have a

fundamental role to play in the compressibility of a rock mass even when it is

cut by joints. The normal stiffness of a joint is influenced by the stress-strain

properties of the rock forming the joint walls. A joint with a given surface

roughness and degree of contact in a stiff rock will be more stiff than the

same joint in a more compressible rock. In chalk the stiffness of the intact

rock is related to porosity. The porosity of chalk can vary over a relatively

wide range, hence the intact stiffness will also vary (see section 2.1). This

leads to the hypothesis that the compressibility of high porosity chalk in the

mass is greater than that of a rock mass comprising low porosity chalk. In

reality the situation is complicated by the fact that the porosity of the intact r

chalk is likely to influence the nature of the fractures developed in terms of

frequency, persistence, aperture and surface topography. The interaction of

these factors could give rise to the case where the compressibility of the low

porosity chalk is greater than that of a high porosity chalk.

When a porous rock such as chalk is subject to unconfined compression it

will fail at a given stress leveL If, however, the rock is confined it will not fail,

but a point will be reached when the skeleton of bonded particles collapses.

The stress at which collapse occurs will be dependant upon the bond strength

and to some extent the porosity (see section 2.1). At the points of contact

across a joint, high stresses can be generated even at relatively low applied

foundation loads. These contact stresses may be sufficiently high to cause

either failure or pore structure collapse of the asperities in contact with each

65

other. This is likely to bring about a change in the gradient of the load­

deformation curve for the rock mass. However the situation is complicated by

the scale of the surface roughness of the joints. When the collapse of the

asperities occurs the associated deformation will bring more asperities into

contact or simply increase the area of contact across the initial asperities. As

discussed earlier this is a function of whether the joint is mated or not. The

result may possibly be progressive collapse of the pore structure at points of

contact giving rise to a gradual change in gradient of the load-deformation

curve for the rock mass.

STRESS DISTRmUTIONS IN FRACTURED ROCK

In the case of persistent horizontal and inpersistent vertical joints the stress

bulb associated with an foundation loading will be elongated vertically in

comparison with that in an elastic continuum (Fig. 2.2/16b) due to the stress

transfer across the horizontal joints together with some lateral flaring due to

the inpersistent nature of the vertical joints causing load shedding. The

general shape of the stress distribution in this joint system has been derived

from model tests (Gazievand Erlikham, 1971 and Knox and Mok, 1985).

Where the vertical joints are persistent the rock mass will act like a series of

discrete columns and hence the stress bulb is likely to be even narrower and

extend to greater depths than in the case discussed above. The general shape

of this stress bulb has been investigated using model tests (Gaziev and

Erlikham, 1971) and is shown in Fig.2.2/27d. The degree to which the bulb

extends outside the loaded area depends upon the aperture of the vertical

joints and the stress level. If the vertical joints are tight there will be load

shedding across the vertical joints and the bulb may extend outside the

loaded area by an amount approaching that for the elastic continuum (Fig.

2.2/16a). If the joints are open with little contact across them it is likely that

the bulb will remain within the loaded area. However the vertical columns

may deform laterally with increasing stress level and local yield such that load

66

shedding across the vertical joints is then possible. It is therefore likely that

the shape of the stress bulb will change as the foundation load increases.

If the joints are inclined the stress bulb beneath the loaded area will become

distorted and asymmetrical. The stress bulbs derived from model tests and

mathematical models for different joint orientations are shown in Fig.

2.2/16c.

The model tests on jointed rock masses all show that the stress distribution

beneath the loaded area is significantly different to that based on an elastic

continuum. In a real rock mass such as the chalk the situation is further

complicated by the fact the style of jointing (orientation, persistence and

aperture) often changes with depth. It is arguable therefore whether the

stress at a given point below a loaded area can be predicted with any

accuracy.

Load-Deformation Behaviour of Fractured Chalk

On the basis of the forgoing discussion a concave load-deformation

relationship for chalk dominated by sub-horizontal bedding discontinuities

and sub-vertical joints. However, the pressure-settlement curves for plate

loading tests on weathered near surface chalk were generally convex (Lake

and Simons, 1975, Hodges, 1976, Marsland and Powell, 1981 and Powell et al.

1990). The characteristic shape of the pressure-settlement curve led Burland

and Lord (1970) to idealise it as a bi-linear curve (Fig. 2.2/17). This

idealisation permits the pressure-settlement behaviour of the rock mass to be

described using five parameters:

E j - initial modulus

Ey - post-yield modulus

qe - yield bearing pressure (based on the onset of yield)

qy - yield bearing pressure (based on the establishment of Ey)

67

Burland and Lord suggested that the change in gradient of the pressure­

settlement curve is associated with yielding. This is based on the results of

865mm diameter plate loading tests at Mundford, Norfolk which displayed

little irrecovable deformation in unload-reload cycles prior to the change in

gradient and significant irrecoverable deformation after the change in

gradient. Little is understood concerning the mechanism by which the chalk

yields. Lord (1990) suggests that this yield is associated with slip and possibly

breakage between blocks of intact chalk.

Behaviour of Foundations on Chalk

The behaviour of foundations on chalk may be studied from case records

relating to full scale structures which have been monitored. The results of

plate loading tests may also be included were the plate diameter is greater

than 1m, since the scale is approaching that of a full scale foundation. The

literature, however, reveals a very limited number of case records of full scale

foundations or large diameter plate loading tests. A summary of these case

records is given in Table 2.2/2.

Of the 6 case records outlined in Table 2.2/2 it will be seen that the

majority are for foundations on structured chalk. There is only one published

record of a full-scale foundation loading the chalk beyond the yield bearing

pressure qe and only one case where long term settlement observations have

been made in order to study time-dependent deformation. The case records

presented here a diverse in terms of the rock mass conditions, the

instrumentation employed, the assumptions made concerning contact stresses

and stress distribution with depth, the control over rate of loading and

unloading and the duration over which observations were made. As a result it

is impossible with such a limited database to identify patterns of behaviour

that would significantly improve our current level of understanding

concerning the mass compressibility of chalk.

68

Table 2.2/2 Summary of published case records of the behaviour of foundations on chalk..

Location Details Source

Mundford, Norfolk 18.3m diameter full scale tank Ward, Burland & Gallios test. (1968).

Grade IV 1m chalk, Pd = 1.55 Mg/m3

Bury St Edmunds, Suffolk 22.85m diameter rafts for silos. Kee (1974), Burland & Grade IV 1111 chalk P g = 1.35 to Davidson (1976) and

1.45 Mg/m Burland & Bayliss (1990)

Basingstoke, Hampshire 1.98m and 3.35m square footings. Lake & Simons (1975) Grade m chalk, P d = 1.34 to

1.43 Mg/m3

Reading, Berkshire 16.5m and 14.5m square rafts. Burland, Kee and Burford Grade V IV! chalk, inferred from (1975)

SPT tests.

Salisbury, Wiltshire 1.55m Dia. footing loading test. Burland, Hancock & May Grade V! chalk. (1983)

Luton, Bedfordshire 3 * l.5m slab and 1.7lm Powell, Marsland, Longworth diameter pad loading tests. & Butcher (1990)

Grade III chalk.

The case records are presented below in the same order as they are shown in

Table 2.2/2. This order reflects the contribution made to the state-of-the-art

as well as the different rock mass conditions. In all the case records the rock

mass has been classified using the engineering grade classification developed

for the chalk at Mundford, Norfolk by Ward et aI. (1968). This classification

scheme is discussed later and the definitions of the grades are given in Table

2.3/5.

69

22.85m DIAMETER RAFfS FOR SILOS AT BURY ST EDMUNDS , SUFFOLK

Burland and Davidson (1976) describe a study of the settlement of some silos

on Upper Chalk at Bury St Edmunds, Suffolk. Kee (1974) and Burland and

Bayliss (1990) present a more detailed study of the settlement of each silo.

The silo complex comprises four cylindrical silos of 12000 tonne capacity

arranged in line with an elevator tower at one end (see Fig. 2.2/18). The

dead weight of each silo is approximately 4100 tonnes.

Each silo has an internal and external diameter of 20.12m and 20.57m

respectively and is supported on a raft foundation 22.86m in diameter and

1.22m thick bearing directly on the underlying chalk. The silos are spaced at

22.81 m centres and the rafts are separated by joints into independent units.

The rafts, which are reinforced, are 1.22m in thickness. Internally each silo

has a floor supported at a height of 2.54m above the base on short concrete

columns. The floor is independent of the silo wall. The layout of the silo

complex is shown in Fig. 2.2/18.

The geology of the site consists of glacial material up to about 4m thick,

overlying Upper Chalk to great depth. The chalk in this area is sub­

horizontal. The chalk at this site has been inspected in the walls of 965mm

diameter shafts which were augured to depths of 15m and 19m at locations

shown in Fig. 2.2/18. The porosity of intact chalk samples taken from a

borehole a few metres west of the silos was found to be between 45% and

49% to a depth of 26m. Below this depth the porosity steadily reduces to a

value of 39% at 29.5m (Burland and Bayliss, 1990). Such high intact

porosities directly below the silos will give rise to low values of intact stiffness

(see section 2.1) which is likely to influence the compressibility of the rock

mass at this site. Burland and Bayliss (1990) noted from the observations in

the shafts that the bedding planes in the chalk often displayed separation with

pillar-like supports suggesting that there is limited contact across these

discontinuities. A low contact area across the sub-horizontal discontinuities

70

combined with a chalk of relatively low intact stiffness is likely to give rise to

any relatively low normal stiffness for the bedding planes.

The chalk in the shafts was classified using the Mundford grading system

(Ward et al. 1968 see Table 2.3/5 for definitions). Above about 5m (below

ground level) the chalk on the west side of the silos is highly weathered and

is described by Burland and Bayliss (1990) as Grade VI to V whereas to the

east the chalk is of better quality being described as Grade V to IV. Below

5m depth the chalk is of Grade IV and occasionally Grade III in all the

shafts. The base of each raft was located at a depth about 1m below ground

level (Kee, 1974) in structureless chalk ( Grade V /VI, see Fig. 2.2/19).

Structured chalk (Grade IV or ill/IV) occurred about 1m below the

underside of each raft (Fig. 2.2/19).

During November 1973 silo 1 was filled with approximately 10000 tonnes of

sugar (giving 344 kPa bearing pressure) with silo 2 virtually empty. As a

result silo 1 began to tilt towards the south west (Kee, 1974). The settlements

measured in the interior of the silos were about 19mm at the extreme eastern

edge, 47mm at the centre and 87mm at the western edge. It will be seen that

these settlements reflect the change in quality of the chalk from west to east.

The observation shafts were sunk after this event and the visual examination

of the chalk in the shaft walls revealed frequent shattering in the immediate

vicinity of the separation planes usually up to a thickness of 80 - 100 mm

(Burland and Bayliss, 1990). This is clear evidence of crushing and collapse

of the asperities of contact associated with sub-horizontal bedding plane

discontinuities.

Settlement observations of the base of each silo were made using a precise

optical level, BRS settlement bolts, and an invar staff. Vertical deflexion

measurements were made at various depths (up to 30m) beneath the centre

of each silo using circular magnet extensometers inserted in 150mm diameter

boreholes.

71

Each silo was filled to a capacity of between 10000 and 12000 tonnes giving

bearing pressures (including that imposed by the dead weight of the

structure) of between 340 and 400 kPa. Fig. 2.2/20 shows the load-settlement

behaviour of each of the silos on first loading. It will be seen that all the

load-settlement curves are convex, showing the yielding behaviour commonly

observed in results of plate loading tests on the chalk. The measurements on

these silos appear to be the first documented case history of a large-scale

foundation for which yield has been observed. The parameters derived from

measurements of the raft settlements are given in Table 2.2/3.

Table 2:1./3 Parameters derived from average surface settlement

measurement~ (aile Kee, 1974)

Silo

1 2 3 4

Initial modulus

E· 1

(MPa)

785 830 830 800

Post yield modulus

E (MFa)

25 37 52 50

Yield bearing pressure

'Ie (kPa)

225 225 200 200

Yield bearing pressure q (~a)

280 330 270 280

Fig. 2.2/21 shows a typical settlement profile for the base of each silo at

various stages of loading. For sugar loads that up to 5000 tonnes the

settlement profile is typically dish shaped indicating( ignoring the differential

settlement the foundation was behaving in a flexible manner. This loading

intensity lies within the pre-yield portion of the sugar-load-settlement curves

for each silo (Fig. 2.2/20). Within this loading intensity, the settlement of a

silo influences the adjacent one in a similar manner to that predicted using

elastic theory (Boussinesq). However when the sugar-load is greater than

5000 tonnes the settlement profile becomes convex (Fig. 2.2/21) and the silos

appear to act independently (Kee, 1974). This corresponds to the yield and

post yield portions of the sugar-load-settlement curves shown in Fig. 2.2/20. It

was this doming of the foundations that caused the columns supporting the

floor of silos to crack. The change in the deflected shape of the foundations

72

suggests that during the intial stages of loading the sugar within the silos was

applying a uniform stress to the silo floor which gave rise to a relative ley

uniform contact stress. At some point during the loading of each silo the

mass of sugar bagan to arch and cause the live load to be transmitted through

the silo wall. Thus as the load in each silo was increased the maximum

settlement moved from the centre to the edge of each slab.

The relationships between vertical deflexion and depth beneath the centre of

silo No.3 for various average foundation pressures are shown in Fig. 2.2/22.

It will be seen that the greatest settlement occurs at the ground surface and

diminishes with depth. Burland and Bayliss (1990) have used the magnet

extensometer measurements to determine the stress-strain relationships for

elements of the rock mass at various depths below the centre of each silos.

These relationships are shown in Fig. 2.2/23 for silo No.3. It will be seen

that these stress-strain curves all exhibit convex curves. For elements below

Sm depth the gradients of the post yield portion of each curve are generally

similar and steeper than those for elements at depths above Sm. Below Sm

depth a general improvement in the quality of the rock mass was observed in

the shaft wall. Intuitively the stiffness of this chalk should be greater than the

more weathered material above Sm. However the indications from the stress

strain curves shown in Fig. 2.2/23 suggest the reverse. Another surprising

aspect of the relationships shown in Fig. 2.2/23 is the apparent reduction in

yield stress with increasing depth of elements. Of course these stress-strain

plots use the average foundation pressure at the ground surface for the stress

and not the stress at the depths at which the strain determinations were

made.

The stress distribution with depth below the silos is complicated by the ring

loads from the silo walls which tend to dominate the settlement behaviour

after the onset of yield. Fig.2.2/24 shows a comparison between vertical stress

distributions predicted using' Boussinesq and finite elements based on linear

elasticity for silo 2 (Nicoletto, 1979). The soil structure interaction analysis

shows that along the centre line the vertical stress initially increases with

73

depth before starting to decrease at about 12m below foundation level. This

is due to the presence of stress concentration bulbs below the edge of the raft

(see Fig. 2.2/25). These bulbs coalesce at depth such that there is little

difference between the stress distribution predicted by Boussinesq and that

determined from finite elements. This is clearly shown in Figs. 2.2/24 & 25. It

is likely that it is the stress distribution and not the rock mass that is causing

the anomalies in Fig. 2.2/23.

Burland and Bayliss (1990) determine a yield locus for the chalk beneath

silos 3 and 4. To determine the horizontal and vertical stress changes

associated with the foundation loading Burland and Bayliss assumed that the

rock mass was homogeneous and used elastic theory for a rigid circular load.

Clearly this approach is likely to yield erroneous results due to the stress

concentrations discussed above.

Only a few periods of clearly established creep settlements have been

identified (Burland and Bayliss, 1990). In 1976 the load in silo 2 was held

constant for 7 months. During this period the silo settled at a constant rate of

O.4mm per month. Other measurements over shorter time periods gave

similar rates of settlement.

It is likely that the yielding behaviour observed in the chalk beneath each silo

was probably caused by the contact stresses developed at the edge of the rafts

which was much greater than the contact stress based on the known load and

the area of the raft. Hence the reliability of this case record is brought into

question particularly with respect to the post-yield load-settlement behaviour.

The values of post-yield stiffness given by Kee (1974) are likely to be greater

than the true values since they are based on a uniform contact stress

distribution.

74

18.3m DIAMETER TANK LOADING TEST AT MUNDFORD , NORFOLK

The most comprehensive case record of foundation behaviour on chalk is the

Mundford tank loading test. As part of a detailed site investigation for a large

proton accelerator the Building Research Station carried out a loading test

using an 18.3m diameter water filled tank at Mundford, Norfolk. Details of

this test are given by Ward et al. (1968).

The location of the tank test was chosen to be most representative of the

rock mass conditions at the Mundford site. A log of the geology to a depth of

30m below ground level is included in Fig. 2.2/30 with the details of the tank

location and location of the instruments. The tank was 18.3m high and of the

same diameter and was founded 1. 7m below the ground surface on

structureless chalk (Grade V, Ward et al. 1968)

Measurements of vertical deflexions at various depths under and adjacent to

the tank were made using instruments in a series of 0.9m diameter shafts

together with ground surface de flexion measurements made using a number

of precise water level gauges (see Fig. 2.2/26). Details of the instrumentation

are described by Ward et al. (1968). Settlement measurements of the tank

were made by precise optical level observations.

The programme followed in carrying out the full-scale tank test included a

short-term cycle of loading followed by long-term maintained load test. The

short-term test included a loading stage lasting 37 hours followed by a period

of 4.5 days whilst the load was maintained followed by an unloading stage

lasting 23 hours. The average bearing pressure (imposed by the tank) when

the tank was full of water was 180 kPa. This represents the average contact

stress assuming the base of the tank is flexible. It must be pointed out that

the stress distribution beneath the tank is controlled by the discontinuity

75

pattern and hence the vertical stress at a given depth cannot be predicted

accurately using models based on an elastic continuum.

The vertical deflexions at various depths in shafts S 1 to S4 on completion of

the first filling of the tank are shown in Fig. 2.2/27. It will be seen from this

figure that maximum deflexions were measured beneath the centre of the

tank (1.19mm at level 1 to O.15mm at level 6), with significantly smaller

settlements being measured at the edge of the tank (O.26mm at level 1 to

O.05mm at level 6). The settlements at all the other points outside the tank

were smaller than the accuracy of the water level gauges (ie < O.lmm). A

shortening between levels 1 and 6 measured in shafts S3 and S5 was O.03mm

was indicative of the fact that very small deflexions had occurred outside the

loaded area. The deflected shape of the ground surface is shown in Fig.

2.2/27 where it is compared with that predicted from simple elastic theory. It

is evident that the observed deflexions are localised around the loaded area

far more than the simple elastic model suggests. Ward et al. (1968) suggest

that this behaviour is consistent with theoretical predictions by Gibson (1967)

for an elastic material whose stiffness increases with depth. However it is also

consistent with the stress distribution shown in Figs. 2.2/16b for a rock mass

dominated by horizontal and vertical discontinuities. Since the discontinuity

frequency is shown to decrease with depth at Mundford it does not seem

unreasonable to assume that the stiffness of the rock mass increases with

depth. The observed ground surface profile however, is likely to be a result of

the combined effect of increasing stiffness with depth and the pattern.

The relationships between bearing pressure (imposed by the tank) and

vertical deflexions at various depths below the centre and edge of the tank

are shown in Fig. 2.2/28. It will be seen that the curves are slightly convex,

and that above a bearing pressure of approximately 39kPa they become

linear. The yielding behaviour commonly seen in plate loading tests on chalk

is absent here since the maximum bearing pressure imposed by the tank

(180kPa) was too small to cause the chalk to undergo yield. The chalk

beneath the tank did however, exhibit some creep during the period when the

76

load was maintained. This can be seen in Fig. 2.2/29 where the strains

observed beneath the centre of the tank are plotted against time. It will be

seen that the creep is not significant at depths below the centre of the tank

greater than 7m. No creep was observed outside the tank. During unloading

the strains that occurred between 11m and 29m below the tank were fully

recovered whilst those between 4m and 7m were not. Table 2.2/4 shows the

percentages of the immediate strains recovered on unloading (creep effects

apart). Ward et al. (1968) suggest that for all practical purposes the

immediate deformation in all but the structureless grade V are all fully

recoverable.

Table 2.2/4 Percentage of immediate strains recovered on unloading and

values of Young's Modulus (after Ward et al. 1968)

Level (below tank)

(m)

0.00 to 3.81 3.81 to 7.32 7.32 to 10.93

10.93 to 14.59 14.59 to 22.57 22.57 to 28.97

% recoverable strain

(immediate)

75.0 97.3 99.7

100.0 100.0 100.0

Young's Modulus (MPa)

1st loading 2nd loading

363 451 1658 1560 1628 1619 2178 2070 4287 3944 4611 4277

Grade of

chalk

V IV - III

III III - II

II IT

In Table 2.2/4 values 'of Young's modulus for the first and second loading at

various depths have been listed. These values have been estimated from the

immediate vertical strains measured beneath the centre of the tank. However

the stress distribution beneath the tank was determined using simple isotropic

elastic theory. Since it is known that the stress distribution a discontinuous

rock mass cannot be accurately predicted using this simple model (Knox and

Mok, 1985) it is likely that the values of E given in Table 2.2/4 are in error.

The second stage of the tank loading test involved a long-term loading test to

investigate the time-deformation characteristics of the chalk. The time­

deflexion relationships for level 1 relative to level 6 in each of the 5 shafts

77

(see Fig. 2.2/26) for the first 18 months after filling the tank are shown in

Fig. 2.2/30. It will be seen that most of the creep settlement occurs directly

beneath the tank with very little creep observed outside the tank. It should be

noted that the ground adjacent to the edge of the tank experienced a

negative creep. Ward et al. (1968) attribute this unusual behaviour to the fact

that the groundwater table rose approximately 3m over the period during

which these observations were made.

Fig. 2.2/31 shows the time-deflexion relationships for three levels beneath the

tank over a period of about 1 year. It will be seen that the majority of the

creep takes place in the top 7 metres (ie in grades IV and III chalk). The

grade III chalk between levels 2 and 3 showed a small amount of creep which

terminated within 120 days. The Grade IV chalk between levels 1 and 2

showed significant creep which was by no means complete after 1 year. This

result seems to be the only well documented case of creep on chalk and it is

unfortunate that the test had to be terminated after only one year.

On the basis on the Mundford tank test Burland (1975) suggests that the

ratios between long-term and short-term moduli for Grades V, IV, and ill

are 0.2, 0.3, and 0.5 respectively.

3m * 1.5m SLAB AND 1.71m DIAMETER PAD LOADING TESTS AT

LUTON, BEDFORDSHIRE

As part of the investigations for bridge works in the Luton area, large scale

in-situ loading tests were carried out by BRE. The results of these tests are

reported by Powell et al. (1990). The regional dip of the chalk in this area is

0.50 to the south-east. However Sylvester-Bradley and Ford (1968) report

some gentle folding and minor faulting of the chalk in Luton area. Despite

these minor structural features it reasonable to expect sub-horizontal and

sub-vertical discontinuities to dominate the rock mass.

78

A loading test was made on a 3m long and 1.5m wide reinforced concrete

slab constructed in an excavation with an invert level 4m below ground level

in Grade III/IV chalk. The load was applied using the 500 tonne frame

described in Chapter 4 up to the full capacity giving a maximum bearing

pressure of 1100kPa.

The chalk below the level of the slab test comprises blocky chalk with sub­

horizontal joint spacing 30-100mm and apertures of 0-3mm. At a depth of 3m

below the slab test the discontinuities become generally tight and more

widely spaced (Grade II chalk). This improvement in the quality of the chalk

is reflected in the SPT N values and ultimate bearing pressures measured at

various depths (see Fig. 2.2/32).

The pressure-settlement behaviour displayed by the slab loading test is shown

in Fig. 2.2/33. It will be seen from this figure that the chalk showed no clear

signs of yielding up to the maximum applied bearing pressure of 1100kPa.

The loading sequence employed involved a number of unload-reload loops. It

will be seen from Fig. 2.2/33 that the gradient of the pressure settlement

curve for each loading stage becomes less steep during the first two cycles

indicating a decrease in compressibi1jty with increasing bearing pressure. The

pressure-settlement relationship for the first loading stage is concave. This

type of pressure-settlement behaviour is commonly seen in stronger rock

types where normal closure of discontinuities is the dominant deformation

mechanism (see section 2.5). However this is the first documented loading

test where this behaviour has been observed in the chalk. Fig. 2.2/34 shows

the pressure-settlement relationship for the slab test with the creep

settlements removed. It will be seen that the pressure-settlement curve is

concave up to a bearing pressure of 550kPa. At bearing pressures between

550kPa and 800kPa the relationship is more or less linear and above 800kPa

the gradient increases slightly. This slight change in gradient at 800kPa

bearing pressure may indicate the onset of yield.

79

The concave pressure-settlement relationship may be associated with the fact

that the sub-horizontal discontinuities became tight within 3m of the base of

the slab. Such tight fractures are likely to have a significantly higher contact

area than those in more weathered chalk which may display apertures greater

than 3mm. However it may be argued that the initial concave part of the

pressure-settlement curve is associated with disturbance of the base of the

excavation for the slab, since machines were used to make the excavation.

At the end of each loading stage the load was maintained for a period of

time between 70 and 134 hours. Time-dependent settlements of between 0.5

and 2mm were observed during these periods. There does not appear to be

any correlation between the bearing pressure and the amount of time­

dependent settlement. The irrecoverable deformations at the end of each

unl?ading stage are similar in magnitude to the deformation measured during

the periods whilst the load was maintained. This indicates that the immediate

deformations were largely recoverable.

Powell et al. 1990 describe a load test carried out on a 1710mm diameter pad

footing at 2.5m depth. The rock mass at this site is described in Fig. 2.2/35.

The pad bears directly on material classified as Grade m chalk. Powell et al.

noted that the Grade IV /m chalk at the location of this test had a more

open macrofabric than that observed by BRE at Mundford. The

topographical position of the site, in the bottom of the Lea valley, suggests

that this open macrofrabic may have resulted from the chalk being subjected

over a prolonged period to strong lateral flow of groundwater. Such

conditions would result in dissolution of material from the walls of

discontinuities and reduce the contact area across these fractures.

The pressure-settlement behaviour displayed by the pad loading test is shown

in Fig. 2.2/36. It will be seen from this figure that the chalk showed no clear

signs of yielding up to the maximum applied bearing pressure of 600kPa. This

is better seen in Fig. 2.2/37 in which only the loading stage of each unload­

reload cycle have been shown and the settlements associated with the periods

80

of maintained load have been removed. The pressure-settlement relationship

shown in Fig. 2.2/37 is more or less linear unlike the concave relationship

observed for the slab test discussed above. Both tests were performed in

Grade III chalk. However it cannot be argued that the open macrofabric is

responsible for the difference in shape of these pressure-settlement

relationships, since the location of the slab test the chalk was observed to

improve to Grade II within 3m of the base of the slab, whereas the quality of

the chalk was only subject to visual inspection to a depth of 1.5m below the

pad. The difference in the shape of the two curves may be attributed to the

fact that the test level for the pad was carefully prepared by hand. Hence

disturbance to the underlying chalk was minimised.

1.98m AND 3.35m SQUARE FOOTINGS AT BASINGSTOKE,

HAMPSHIRE

Lake and Simons (1975) made settlement observations on a four-storey

reinforced concrete framed structure supported on spread foundations in

chalk in Basingstoke, Hampshire. The chalk at this site is described as being

weak to very weak (ac = 0.4 to 1.0 MPa). At shallow depths the discontinuity

spacing was between 50 and 100mm. Below 2 to 3m depth the spacing

increased up to 200mm and the fractures were generally tight. The bedding at

this site is described as horizontal, suggesting that this area has not been

subjected to intense tectonic activity. It is therefore reasonable to assume that

sub-vertical and sub-horizontal discontinuities dominate the rock mass. It is

not clear from the description of the rock mass given in the paper whether

discontinuity spacings relate to sub-horizontal or sub-vertical discontinuities.

However given that most of the data collected was from boreholes it is likely

that the descriptions relate to sub-horizontal discontinuities.

A simplified foundation plan for the structure is shown in Fig. 2.2/38.

Observations of settlements were continued for 36 weeks, commencing within

a few days of the stub-column shuttering being struck. Settlement

81

observations were made using a precise optical level on six internal columns

with bases 3.35 m square and four external columns with bases 1.98 m square.

The larger bases were located 1.5 m to 3.0 m below the ground floor slab

level and the smaller bases 1 m below this level. Most of the footings were

therefore placed in the chalk with the 200mm fracture spacing, which is

described as being Grade III to II on the basis of fracture spacing and SPT N

values. The design bearing pressure was 430 kPa and that due to the

structural loading was calculated to be 365 kPa, which is the value adopted

for determining the values of Young's modulus.

The measured settlement for the 3.35 m and 1.98 m square bases are shown

in Fig. 2.2/39. The average settlement at the end of the observation period

was 6.5 mm for the large bases and 4.0 mm for the small bases. About 50%

of the total average settlement was recorded within 6 weeks, before the first

floor slab had been cast. Very little settlement was observed after 10 weeks.

Lake and Simons suggest that a large proportion of the initial settlement was

due to the compression of loosened chalk which was disturbed during

excavation. The values of Young's modulus calculated on the basis of the

total average settlements after 36 weeks were 105 and 95 MPa for the 1.98 m

and 3.35 m bases respectively. These values are very low compared with that

derived from the Mundford tank test for Grade III and II chalk. The dry

density of the chalk at Basingstoke is lower than that at Mundford (see Table

2.2/2) indicating a higher porosity and hence a lower intact stiffness. The

plate loading tests carried out at this site showed the typical yielding

behaviour with yield stresses (qy) between 363 kPa for a 0.91m diameter plate

to 527 kPa for a O.3m diameter plate. There is clearly a trend of decreasing

yield stress with increasing plate diameter. The relationship between yield

stress and plate diameter for this site is not linear and at plate diameters

greater than 0.61m the yield stress in much less sensitive to increasing plate

diameters. On the basis of the plate tests it is likely that the yield stress

applicable to the bases is about 360 kPa. Since the structural loading placed

on these bases is considered to give rise to a bearing pressure of 365 kPa it is

likely that the chalk beneath the bases is undergoing yield. It may be argued

82

however that since the plate loading tests were carried out at a depth of 1.5

m the chalk was more closely fractured than at the depth of the foundations.

Hence the yield stress at the foundation level may be greater than that

determined from the plate loading tests. The initial modulus (Ej ) determined

from the 0.91m diameter plate was 443 MPa and the post yield modulus (Ey)

was 15 MPa. The modulus determined from the observed settlements at an

assumed bearing pressure of 365 kPa is intermediate between these two

values. It i~_!ikely that the disturbance caused during excavation together with

the onset of yield beneath the bases contributed to the low mass stiffness

values determined from the observed settlements.

16.5m AND 14.5m SQUARE RAFTS AT READING, BERKSHIRE

Burland et al. (1975) made settlement observations of a five-storey building

founded in soft weathered chalk at Reading, Berkshire during construction.

The building was of reinforced concrete frame construction with brick

cladding. The foundations comprised three discrete rafts with joints between

them. The rafts are 1m thick and were cast directly onto the surface of the

chalk. A plan of the raft foundation is shown in Fig. 2.2/40.

The site investigation was carried out by sinking three 150mm diameter

boreholes by means of a light cable percussion rig, and digging three trial

pits. The positions of the boreholes and trial pits are shown in Fig. 2.2/40. "It

was not possible to characterise the chalk by visual inspection at this site due

to depth of cover and the high water table. The chalk was therefore

characterised using the Standard Penetration Test (SPT) and the approximate

correlation between the Mundford grading and SPT 'N' value given by

Wakeling (1970). The ground conditions beneath the site are shown in Fig.

2.2/41. No dry density values are given by Burland et al. (1975) for the intact

chalk. This makes it difficult to make meaningful comparisons with

foundation behaviour observed at other sites. There are some similarities

with the ground conditions beneath the Mundford tank in that the chalk

beneath the raft is essentially structureless. However at the Reading site the

83

thickness of structureless chalk is much greater and the groundwater table is

much higher.

Settlement observations were made of a number of points on the raft (see

Fig. 2.2/41) using a precise optical level and an invar staff. The relationship

between settlement and time for some of the levelling stations in Block A is

shown in Fig. 2.2/42. The foundations of the building were constructed in

August 1972 and by May 1973 the building was substantially complete. Most

of the points show between O.5mm and O.lmm settlement at very low load.

This may be due to the bedding of the raft into chalk disturbed by

excavation. During the initial stages of construction the rate of settlement

with time is small and is considered to be elastic by Burland et al.. Towards

the end of construction the rate of settlement is significantly higher. It is not

clear whether this is a result of a reduction in stiffness as a result of yielding

or if it associated with the onset of creep.

Values of E based on the applied stress (57 kPa) and the average settlements

(corrected for settlement associated with bedding of raft) at the end of

construction for Blocks A and Bare 113 MPa and 164 MPa respectively. At

Mundford the value of E for the structureless chalk beneath the tank

immediately after loading was 360 MPa. Over a period of one year under a

maintained load creep settlements confined to the more weathered chalk

resulted in a reduction in stiffness. The ratios between long-term and short­

term moduli for Grade V chalk is 0.2 for the Mundford tank test (Burland,

1975). Hence the long-term value of E for Grade V chalk at Mundford was

72 MPa. At Reading the major part of the loading was applied over a period

of about six months. Hence a significant reduction in stiffness with time may

be expected. The lack of agreement between the stiffness values determined

for Mundford and Reading may result from the different time periods

involved, the different groundwater levels and differences in intact

mechanical properties of the chalk at each site.

84

1.55m DIAMETER IN-SITU LOADING TEST AT SALISBURY,

WILTSHIRE

Burland et al. (1983) describe the results of a large in-situ loading test on

highly weathered chalk at Salisbury, Wiltshire. The test was carried out to

assess the load-settlement characteristics of the foundations for a cable

chamber which forms part of a telephone exchange. The telephone exchange

is a four-storey building of reinforced flat slab construction. Most of the

structure is founded on pad footings bearing on gravel. The cable chamber

runs along one edge of the building. The base of the cable chamber

comprises a reinforced concrete raft 9m wide and 47m long, bearing on

highly weathered structureless chalk some 6.6m below original ground level.

The gross bearing pressure imposed by the cable chamber was approximately

185kPa Excavation for the cable chamber took place within a cofferdam due

high ground water table which was about 2m below original ground level. The

chalk exposed in the base of the cofferdam was described as very soft,

consisting of small angular pieces of chalk in a silt like putty chalk matrix.

The nature of the chalk and the high water table made excavation difficult

without causing significant disturbance to the formation level. The degree of

disturbance and the general condition of the chalk below the excavation was

investigated using cone penetrometer tests. The poor quality of the chalk at

formation level made it necessary to carry out an in-situ loading test to assess

the load-settlement behaviour of the rock mass.

The geology of the site comprises 2 to 3m of medium dense gravel overlying

4 to 5m of soft structureless chalk, containing small lumps of intact chalk

(Grade VI chalk, see section 2.7/5 for definitions). This is underlain by chalk

which has been classified on the Mundford scale as Grade V at the top

improving to Grade III with depth on the basis of measured SPT 'N' values

(Wake ling, 1970).

Fig. 2.2/43 show a cross-section through the load test apparatus. The loading

test was carried out in a 2m square pit l.5m deep, excavated underwater in

85

the base of the larger cofferdam excavation. A loading column 1.55m in

diameter was constructed using thrust-bore rings. A concrete plug was

formed in the first ring and measures were taken to ensure adequate load

transfer between the ring and the plug. A 153mm dia. tube was cast into the

plug to permit measurements of the settlement of the footing and to allow

measurements to be made of ground movements below the footing. These

measurements were made using a 53mm dia. steel tube and a 25mm dia.

steel rod which were grouted into the chalk at a depths of 1.19m and 2.13m

below the base of the pit respectively (see Fig. 2.2/43). It should be noted

from Fig. 2.2/43 that the loading column was not constructed vertical.

The column was loaded using kentledge in 4 tonne increments up to a load

of 40 tonnes which was held for 24 hours. The load was then increased again

in an attempt to reach 60 tonnes. However sever tilting of the column caused

the test to be terminated at a load of 51 tonnes. During the test the 25mm

dia. steel rod became jammed and hence there were no measurements of

settlement at a depth of 2.13m below the footing.

The load-settlement relationship for the footing is shown in Fig. 2.2/44. The r

application of the first loading increment results in a slight heave of the

measuring point below the footing and only a small settlement of the footing

itself. The second loading increment results in a much greater settlement of

the footing together with the settlement of the sub-footing measurement

point. This behaviour is somewhat unusual, but it may have occurred as a

result of the loading column not being vertical. Lateral movements of the

loading column during the test are not reported by Burland et al. (1983). For

the loads between 8 and 36 tonnes the load-settlement curves are

approximately linear giving Young's modulus values of 4.4MPa and 5.5MPa

for the footing a the point 1.19m below the footing respectively (Poisson'S

ratio was assumed to be zero in the determination of these values (Burland

et al. 1983». A slight change in gradient occurs above 36 tonnes (191 kPa

bearing pressure). This may indicate the onset of yield. The large settlement

86

recorded when the load reached 51 tonnes is likely to be associated with

bearing capacity failure.

The values of Young's modulus derived from the loading test are very small

compared to those derived from full-scale foundations and loading tests at

other sites. It should be remembered that this loading test was carried out in

structureless Grade VI chalk, below the watertable in an excavation made

underwater. It is difficult to assess which of these factors will contribute most

to the observed load-settlement behaviour. Clearly the base of the pit must

have been disturbed during excavation underwater This alone would

contribute to a low modulus. However it is surprising that such disturbance

was not reflected in large settlements occurring during the first loading

increment. The nature of the material on which the loading tests was

performed was such that it is more likely to behave as a soil than a rock. The

modulus values obtained are certainly closer to those that would be expected

in a soil, rather than a rock.

The time dependent settlement observed under constant load during the

loading test was very small. Fig. 2.2/45 shows the time-settlement

observations on the cable chamber during and after the construction of the

telephone exchange. The most striking feature is that most of the settlement

took place during the construction of the raft and thereafter the settlements

were small and very uniform ranging from 3 to 5mm.

Normalised Load-Settlement Behaviour

The load settlement behaviour for chalk has been observed in both large

scale loading tests and full scale foundations. It is valuable to be able to

compare the observed behaviour of foundations and such tests. In order to

make such a comparison it is necessary to normalise the data to bring it to a

common base. Powell et al. (1990) suggest the application of a scaling factor

derived by equating the expressions used to determine the elastic modulus

from a rigid circular plate at a given depth below ground level and that for a

87

rectangular foundation at the surface. They point out that such a scaling

factor can be overestimated where the modulus increases uniformly with

depth from zero at the surface and in such cases only provides an upper

bound value. Lord (1990) extended this approach to make allowance for the

rigidity of the foundation. It is, however, unwise to include rigidity in the

scaling factor since there is uncertainty as to whether certain foundations are

to be treated as flexible or rigid. Implicit in both Powell et al. and Lord's

scaling factor is an equivalent plate at depth. It seems more reasonable to

consider as a basis for comparison a circular plate founded at the surface.

Furthermore if average stresses and average settlements are considered, the

error introduced by not making allowance for rigidity will be small. The

writer therefore proposes using an equivalent flexible circular plate founded

at the surface and using only average settlements as a basis for deriving a

scaling factor. The settlement of a foundation can therefore be related to the

settlement of an equivalent plate by means of the expression:

Where:

pr/Bc = S(pp/Dp)

Pc = settlement of foundation

Be = width of foundation

Pp = settlement of equivalent plate

Dp = diameter of equivalent plate

S = scaling factor

The appropriate scaling factors have been applied to the average observed

settlements in the above cases and the resulting load settlement behaviour is

shown in Fig. 2.2/46. Fig. 2.2/46b shows the load settlement behaviour up to

a settlement ratio of 0.4% and hence includes the post-yield behaviour. The

maximum settlement ratio used in Fig. 2.2/46a is 0.1%, allowing the initial

pre-yield behaviour to be examined in detail. Since there is only one

documented case where the chalk has been loaded beyond yield it is difficult

to draw many conclusions from Fig. 2.2/46b.

The typical range of initial modulus values is shown by the envelope in

Figure 2.2/46a (200 to 1000 MPa). Although this range appears large it

88

should be noted that generally pre-yield settlements are small with settlement

ratios less than 0.05 % at average applied foundation pressures up to 200 kPa.

Thus even for foundations 20m in diameter, on grade IV /m chalk,

settlements below this stress level will not exceed lOmm.

Plate Loading Tests on Chalk

Plate loading tests provide a high degree of similitude with the full-scale

foundation. As a result this form of in-situ test is useful for studying the mass

compressibilty behaviour of fractured rocks such as chalk. However the

rigidity and geometry of the plate are such that different elements of ground

are subjected to different stress paths resulting in problems extrapolating

from test-scale to full-scale. This is particularly relevant when the plate

diameter is small compared with the fracture spacing. In general the plate

diameter must be at least six times the average sub-horizontal discontinuity

spacing in chalk (Lake and Simons, 1975). For most near surface weathered

chalks a plate diameter of 865mm is suitable. The use of such a plate

diameter requires a large test rig together with large amounts of kentledge or

the use of tension piles, all of which makes this form of in-situ test

prohibitively expensive for routine investigations. However despite the

expense the literature reveals more results of plate loading tests on chalk

than observations of full-scale foundation behaviour. Table 2.2/5 summarises

the results of plate loading tests on chalk found in the literature.

It will be seen from Table 2.2/5 that in many cases the post-yield modulus Ey

was not established. In some cases this was due to the fact that the yield

stress was greater than the capacity of the plate testing equipment (eg Grade

II chalk at Mundford) or the test was not designed to determine this

parameter (eg constant rate of penetration (C.R.P.) tests at Welford, Theale

and Norwich).

The most comprehensive plate loading test data for the chalk was from tests

carried out at Mundford, Norfolk (Ward et al. 1968 and Burland and Lord,

89

I { y:> II nIl I!".} )1 Ji I{'-..,\ 2-.', c::.->L';VL>'-.J LJULJU c:::#

Site &

Source

Welford & 140 Theale (Lake & Simons (1970» • Initial loading Imm/min " Reloading 2 mm/min

Norwich (Butler & Lonl (1970»

(C.R.P.)

Loading at Imm1min

Bumg.toke 305 (Lake & Simons (1975»

Otterboume 152 (Hodges (1976»

Whitchurch 152 (Hodges (1976»

Brighton (Lonl & Davies (1979»

Purf\eet (Palmer (1966) & (1976»

Geulhemmer 400 N etherlanda 280 (Nienhui. & 180 Price (1990» 120 Secant moduli at 600 kPa

Mundford 865 (Warc!, Burland & Ga11io. (1968), Burland & Lord (1970»

Luton -Site C (powell et al (1990»

Luton -Site 0 (powell et al (1990»

P1ate Depth size (mm) (m)

12.1 (C.R.P.) 19.9

7.7 13.8 19.6 10.1

14.5 19.2

4.7

140 5.5 7.6 9.1

10.7 12.2 2.3 3.9 5.4 6.9 8.4

10.4 12.2

1.5 III 610 1.5 910 1.5

1.5 V 305 1.5 762 1.5

1.5 II 254 1.5 762 1.5

450aq Horr.

445 8.7

Horr. II II II II

80 50 30

3.4 V 3.2

6.0 IV 5.4 2.6 7.8

11.4 7.4 9.0

10.9 12.9 13.6 16.6 19.5 16.6 5.6 8.1

865 2.0 (C.R.P.) 2.0 4.0 III

4.0

865 2.0 (C.R.P.) 3.0 3.2 IVnIl

6.6 6.8

10.5 15.0

Grade Eo E,. CIa q. q,., 10%0

(MPa) (MPa) (kPa) (kPa) (MPa)

72'/1 OS' 9.3 34'/176' 4.3 9'/40" 2.6

25' 6.7 107' >16 83'm6' >16

86' 15.5 121'/402" >16

21'

57 2.9 9 1.6

23 1.7 22 1.4

20 2.1 9 1.8

45 1.3 26 1.2 33 1.9 49 18 2.2 19

690 24 350 475 >2.5 m 312 22 214 320 >1.0 III 443 19 214 290 >0.6

1.5 V 10 2 536 1.5 V 5 >0.4

1640 1.5 II 33 4 680 2.8 II 24 >0.5

II 2360 4950 II 3900 7500

5.0

510 910 llOO 1460 II 460 II 620 II 260

3690 37 200 500 IV 2040 43 200 600 2860 42 200 600 III 3400 194 400 650 m 3690 136 450 550 III 4610 146 350 600 IIIIII 1900 II 2380 II 1960-2240 II 2600 II 2380-2720 II 2940-3920 II 1900-2090 II 1960-2940 II/I 6860 I 9210 I 9500

IV 340 1.70 IV 119 10 45 360 1.65

7 320 1.66 III 120 16 45 230 1.80

IV 77 3.4 IVffiI 197 4.2 234 510

III 128 5.2 m 600 200 520 II 307 7.1 II 1600 6.4

90

1970). At this site 865mm diameter plate loading tests were carried out at

different depths in 0.9m diameter shafts. This enabled stiffness parameters to

be measured for the complete weathering profile. The rock mass at this site

was described in detail down to the level of the water table using a number

of large diameter (0.76m) augered boreholes. The engineering grade

classification that is used to describe the condition of the rock mass at the

sites considered in the case histories and in Table 2.2/5 was developed from

these detailed descriptions (see Table 2.3/5). The most notable feature of

the results of the Mundford plate loading tests was the very high values of

initial modulus for chalk of Grade III and IV. The values are greater than the

values of stiffness obtained from tests on Grade II chalk at Mundford and

significantly greater than the stiffness derived from plate loading tests in

Grades III and IV chalk at other sites (eg Basingstoke, Lake and Simons,

1975). Burland and Lord (1970) suggest that the high values of Ei obtained at

Mundford are associated with adhesion between the plate and the walls of

the auger hole due to the plaster of Paris squeezed up during bedding of the

plate.

The plate loading tests reported by Lake and Simons (1970) between Welford

and Theale on the M4 Motorway were carried out using a 140mm diameter

plate loaded at a rate so as to produce a constant rate of penetration of

1mm/min. Such a rate of penetration renders it impossible to determine Ei

and qy with any accuracy. The ultimate bearing capacity measured at a

penetration of 10% of the plate diameter varied between 3MPa and in excess

of 16MPa indicating a wide variation in strength. The dry density of the chalk

at this site varied between 1.3 and 1.4Mg/m3 indicating that this chalk has a

high porosity (48 to 52%) and hence would be expected to have a low

strength. Constant rate of penetration plate loading tests carried out in

Norwich using the same rate of penetration and the same plate diameter

(Butler and Lord, 1970) give a lower ultimate strength since the chalk in

Norwich generally has a higher dry density than 1.3Mg/m3.

91

Lake and Simons (1975) carried out a series of plate loading tests using a

variety of plate diameters, on the chalk at Basingstoke. The results shown in

Fig. 2.2/47 indicate that the observed behaviour is sensitive to the plate

diameter, which raises the question of the validity of extrapolating data

acquired from small diameter plate tests to full scale foundations. The

sensitivity of the observed behaviour to plate diameter stems from the

inability of small diameter plates to test a representative volume of the rock

mass. The discontinuities in the chalk at the Basingstoke site were

predominantly horizontal and vertical and the average spacing of the

horizontal joints was 50 to 100mm at shallow depths and up to 200mm below

2m. The results shown in Fig. 2.2/47 indicate that a reduction in sensitivity to

plate diameter when the plate diameter exceeds 5 times the average joint

spacing (ie 5 * 100mm). Indeed this rule of thumb has been adopted by

BS1377:1990 part 9. Despite the diameter of the plates used by Lake and

Simons the results of all the tests displayed convex pressure settlement

curves conforming to the Burland and Lord model.

Hodges (1976) carried out a sreies of plate loading tests on chalk at

Whitchurch and Otterboume in Hampshire using three different plate

diameters (152mm, 254mm and 762mm). The chalk was classified using the

Mundford grading system by direct visual examination. The chalk at

Whitchurch and Otterboume was classified as Grade IT and V respectively.

The pressure-settlement relationships are shown in Fig. 2.2/48. It will be seen

from this figure that these results show a similar trend to that observed by

Lake and Simons (1975). It appears that for a wide range of discontinuity

spacings and apertures there is a trend of decreasing initial modulus (Ei) and

increasing yield bearing pressure (qe and qy) with increasing plate diameter.

Nienhuis and Price (1990) report the result of a series of horizontal plate

tests carried out in Maastrichtian chalk in the Geulhemmer mine in Holland.

These tests are of particular interest since although the chalk is coarser

grained (medium grained calcarenite, mostly made up of foraminifera about

2.4mm in diameter and about 1mm thick. Crinoid ossicles about 1mm in

92

diameter and shell fragments are also present) than most chalks found in the

U.K it has a high porosity (40 -50%), joints are so infrequent that they may

be individually mapped and bedding is seldom seen clearly. In such a

material the plate tests reflect the behaviour of the rock material due to the

lack of discontinuities. Nienhuis and Price performed tests with plates ranging

in diameter from 30mm up to 400mm. Fig. 2.2/49 show the influence of

plate diameter on modulus. These results tend to show higher values for the

180mm and 280mm diameter plates but the difference in value between these

and the results from the larger and smaller plates is relatively small. This

leads to the conclusion that the modulus values obtained are independent of

plate diameter. The variations in modulus may be attributed to

inhomogeneity within the rock mass. These results when considered alongside

those of Lake and Simons (1975) highlight the influence discontinuities have

on the measurement of rock mass compressibility.

Most plate loading tests are designed to provide information about the short

term behaviour of the rock mass. As a result there is a scarcity of creep data

for the chalk. Burland and Lord (1970) in an attempt to overcome this

problem devised a simple method of comparing the creep results from plate

tests with long-term settlement observations of full-scale foundations. This ,

involved expressing the rate of creep as a proportion of the immediate

settlement in the following manner:

R = Settlement per cycle of log time *100% Observed total immediate settlement

In Fig. 2.2/50 the values of the creep ratio R from plate tests and the

Mundford tank test have been plotted against bearing pressure. For bearing

pressures above that of yield (qe) the R values are clustered relatively tightly

in the range 10 to 15%. For low bearing pressures (at or below qe) the results

are very scattered and the values of R are considerably higher (> 30%). This

93

is possibly the result of different creep laws being associated with pre and

post-yield behaviour.

The creep ratio R assumes the relationship between creep settlement and the

logarithm of time is linear. This relationship has been observed in both short

and long term maintained loading tests (Burland and Lord, 1970, Powell,

1990). However, the immediate settlement required in the calculation of R is

less reliable, since the term 'immediate' has not been defined by Burland and

Lord (1970). Different interpretations of this term can lead to significant

variations in the value of R. Ideally the immediate settlement used to

determine R should be based on creep rate.

The creep measured in tests on grade ill/IV chalk for a borehole and a

trench based plate loading test are plotted as creep rate against log time in

Fig. 2.2/51. It will be seen from Fig. 2.2/51 that the creep rate reduces

rapidly achieving rates less than those proposed by Burland and Lord (1970),

for the rate at which the next increment of load could be applied, in 500 -

1100 minutes. The time to achieve this rate is seen to increase with increasing

stress level. For stress levels below yield the curves all show similar shapes

and fall close to each other. Once the yield stress (qe) is exceeded however,

the creep rates are seen to increse significantly while maintaining the same

basic shape of curve. It will be seen from Fig. 2.2/51 the curves for the

highest bearing pressures in each test the Burland and Lord rate is not

achieved even after 6000 minutes.

Powell (1990) suggests that creep may be of little importance at low stress

levels (ie < qe) with most of it being built out during construction. However

it is clear that at stress levels approaching and exceeding qe creep settlements

can become significant.

94

Summary

The following points arise from this study of the mass compressibility of

chalk:

• The introduction of discontinuities into any rock mass will significantly

increase the compressibility.

• The principal factors affecting the normal stiffness of discontinuities in

rock include:-

contact area;

stress-strain characteristics of the intact rock;

presence of infill material

and the magnitude of the applied load

The most important of these is contact area. This is related to the

surface topography of the discontinuity walls and the degree to which

adjacent surfaces are correllated. Contact area is related to the

magnitude of the applied load and the stress-strain characteristics of

the intact rock. As the applied load is increased the contact area

changes in a manner controlled by the mechanical properties of the

rock forming the asperities of contact.

In chalk the contact area across sub-horizontal discontinuities has been

reduced through dissolution and may be less than 10% of the total

plan area of these fractures.

• The yielding and collapse behaviour of intact chalk may bring about

similar behaviour in discontinuities due to high stresses generated at

the asperities of contact.

95

• The compressibility behaviour of a fractured rock mass is controlled

by:

the orientation of discontinuities relative to the direction of

applied load;

the number of discontinuity sets;

the spacing of discontinuities relative to the dimensions of the

loaded area;

the stress-strain characteristics of the intact rock

and the magnitude of the applied load.

• The available literature therefore suggests that, for a rock mass with

joints parallel and normal to the direction of loading

• the details of contact geometry, and the yield strength of joint

walls normal to the direction of loading, will dominate the

stiffness of the rock mass

• stiffness will increase with increasing load.

Stiffness can be expected to decrease with increasing frequency of the

joint set normal to the direction of loading, but this may well be a

secondary effect if there is significant spatial variation of contact area

and aperture of the joints.

• The stress distribution in rock below a loaded area is controlled to a

large extent by the discontinuity geometry and hence cannot be

predicted accurately using theories based on continuum mechanics.

• The mass compressibility of the English Chalk shows a behaviour in

contrast to that to be expected from the general rock mechanics

literature. Yielding during normal closure of bedding discontinuities

may possibly initially lead to an increase in stiffness, although available

data suggest that this does not occur. At relatively low applied

96

stresses, in near-surface fractured Chalks, yielding commences and

there is a large decrease in stiffness with increasing load.

• Burland and Lord have proposed an bi-linear idealisation for the load­

settlement behaviour of fractured chalk. Using this model the mass

compressibilty behaviour may be described using the following

parameters:

Ei - initial modulus

Ey - post-yield modulus

qe - yield bearing pressure (based on the onset of yield)

qy - yield bearing pressure (based on the establishment of Ey)

• Only six case records on the behaviour of foundations on chalk exist.

Of these only four deal with foundations on structured chalk. The

reliability of some of these case records is brought into question

because of the assumptions used in determination of contact stresses

and the stress distribution beneath the loaded area (eg Bury St

Edmunds sugar silos and the Mundford tank loading test).

• Generally pre-yield settlements are small with settlement ratios less

than 0.05% at average applied foundation pressures up to 200 kPa.

• Little is known about the post-yield behaviour of foundations on chalk

since there is only one documented case.

• Little is known about the time-settlement behaviour of foundations on

chalk.

• Little is known about the mechanisms causing the rock mass to yield.

• Little is known about the influence the intact mechanical properties of

chalk has on the mass compressibility behaviour.

97

• Most published results of plate loading tests are unreliable since the

plate diameter used was too small in relation to the discontinuity

spacing. In general the plate diameter should be at least five times the

average spacing of the sub-horizontal discontinuities.

98

o

-E E -.. ::I: 0.1 c CI) en c cu .c (J

0.2

Fig. 2.2/1

Normal stress (MPa) o 20 40 Solid rock

t

o . nJ

Interlocked joint \.

" " " "- ...

" ... "-°nt .... " "- ...

..... ..... '- ...

...... ..... t ..... ..... ....... ........ ....

\ aperture = 0.20mm \

\ O. \ nJ

\

\ \ Mismatched joint \ \

"- " ... \ ... \ ... ..... \.

..... \. ..... \. .....

°nt ..... ...

..... \. "'0.\..

6 = 6 - 6 aperture = 0.35mm nj nt nr

Concave normal stress-deformation behaviour of rock joints (after Bandis et al. 1983).

99

Normal stress (MPa) o 10 20 30 40 so

o I

~\~:::::::::.:::::::;::::::::: ................ . 0.02 ___ ::::: _____ _ -------=---

3rd cycle

2nd cycle ~

E .! 0.04

CD -~ tn 0.06 o -(,) .. .E 0.08 o ..,

Fig. 2.2/2

Fig. 2.2/3

0.10

0.12

(I) ., CD -;; -as CD .c

UJ

1st cycle

Umestone (bedding)

Measurements of the closure of natural joints under normal stress (after Bandis et al. 1983).

4 i--f ----;:.---~~

Shear displacement

Convex shear stress-displacement behaviour of rock joints illustrating sample size dependence (after Bandis et al. 1981).

100

Fig. 2.2/4

ca ep r-

Mated joint

-- -- -

Non-mated joint

Mated and non-mated joints.

100 .-------------~--------------~------------~

80 D

o D

Chalk

Sandstone (strong)

Sandstone (weak)

" ca 60 t) ca ... c o

00 () 40 ... D o o

o D o

C ep () r­ep

o --

a.. 20

Fig. 2.2/5

D ---.-,----,-

o ~~ ____________ L-____________ ~ ______________ ~

o 5 10 15

Average normal stress (MPa)

Relationship between joint contact area and average normal stress for a variety of rock types including chalk (after Duncan and Hancock, 1966).

101

Normal stress 3MPa

Normal stress 33MPa

0 20 0

-E E -CD ~

~ 0 0.005 0 -()

..... c: .-0 -,

0.010 as 50 CD ~

as 40 ..... () as 30 ..... c: 0 20 ()

..... 10 c:

CD () 0 ~

CD 0 D.. 20

·~-. ..:. ... ~~~.

". ;.. ''-4-~

,;. . . .~ .

. ,- :":.~ .~. "~ '--'.-~. '

f < . ",. ',

Normal stress 85MPa

Specimen E30 Approx. scale

Normal stress (MPa) 40 60

40 60

Normal stress (MPa)

0.4mm

80

E30

E32

E30

80

(a)

(b)

Fig. 2.2/6 Relationship between joint contact area and average normal stress for natural fractures in quartz monzonite (after Pyrak-Nolte et al. 1987, 1990).

102

Fig. 2~/7

p Type A p

---

Settlement

p Type B p

Settlement

TypeC p

, Settlement

Constrasting load-deformation behaviour for rock masses with different magnitudes of joint shear (S) and normal deformation (N) components (after

Barton, 1986).

103

..-. E E '-"' ... c -­CI)

E CI) () C'G Q. o C

..-. E E '-"' ... C CI)

E CI) -:= CI)

en

Fig. 2.2/8

o o

0.25

0.50

0.75

1.00

o o

0.25

0.50

0.75

1.00

1.25

1.50

1

p

Load (kN) 2

p

3 4

Elastic response

Hydrostatic loading

p (a)

Bearing pressure (kPa) 500 1000 • 1500

(b)

(a) Deformational response where the joint sets are perpendicular to the imposed principal loads (after Chappell, 1979). (b) Results of a plate loading test on sandstone with predominantly sub-horizontal and sub-vertical joint sets (after Hobbs, 1973).

104

Fig. 2.2/9

Pv

.................................. :.:.:.:.:.:.:-:.:.:.:.: :.:.:.:.:.:.:-:.:.;.:.:.:.:.:

:-:-:.:.:.;.;.:-:.;.;.;.: ............................... :-:-:.;.:.;.;.;.: .

. :.:-:-:.:.: ;.;.;.;.;-:.;.:-:.;.:.;,' .... . .::::::'

P':';"'Z:':':':',Z":-:':-:B' ::::::::::::::::==:':B':':':':B.:.:.:-:.q ..... ________ Joints

Pv (a)

:

~i ~ : :

.' ;::: : : :::

:: ::: : : : : .'

.'. ~lj

: ::: ~~ : : : .'

: : : :

:: : :

: .'

: : .' j ~

Pv (b)

Pv (c)

Simple models for studying the influence of joint spacing on mass compressibility.

105

1

0.8

Em 0.6

EI 0.4

0.2

0

Fig. 2.2/10

1

0.8

Em 0.6

E, 0.4

0.2

0

Fig. 2.2/11

EJ E = 0.5

I

EJ E = 0.05

I

EJ E = 0.0001

I EJ E = 0.005

I I Aperture = 1.0mm

0 4 8 12 16 20 24 28

No. of joints per metre

Variation in the ratio of mass stitTness to intact stitTness with fracture frequency for fractures with ditTerent stitTnesses.

Aperture

1 0.005

0 4 8 12 16 20 24 28

No. of joints per metre

Variation in the ratio of mass stitt ness to intact stitTness with fracture frequency for fractures with ditTerent apertures.

106

.-... 0 .., () as .. tJ) tJ) as E

JIll:: () 0 a:

Fig. 2.2/12

.-... 0 .., () as .. tJ) tJ) as E

JIll:: () 0 a:

Fig. 2.2/13

Chalk grades (Ward et al. 1968) Grades I & II Grade III Grade IV 1~---:----------__ ~--------__ ~

0.8

0.6

0.4

0.2

0 0

T1ghtr closed joints

rocompressed joints

4 8 12 16 20 24 28

No. of joints per metre

Variation of rock mass factor (j) with fracture frequency for chalk based on laboratory tests on artificial joints (after Wakeling, 1975).

36.5% contact area ratio

1

0.8

0.6

0.4

0.2

0 0 4 8 12 16 20 24 28

No. of joints per metre (1 If)

Na = 50 Poisson's ratio = 0.25

Variation of rock mass factor (j) with fracture frequency for joints with different initial contact area ratios.

107

'-~ 0 ..., (,) «S ... U) U) «S E ~ (,) 0 a::

Fig. 2.2/14

'-~ o ..., (,) «S ... U) U) «S E ~ (,) o a::

Fig. 2.2/15

O.S

O.S

0.4

0.2

0 0 200 400 SOO SOO 1000 1200 1400

No. of asperities per unit area (Na )

A j = 20%, Poisson's ratio = 0.25

Variation of rock mass factor (j) with joint roughness expressed as the number of asperities per unit area (Na) for rock masses with different joint spacings.

1.0

O.S_

o.sB

0.4

0.2

Killingholme, Lincs [ ::::::::::::::::::::::::::::J Mundford, Norfolk ------ Newmarket, Cambs.

Fractures becoming tighter

0.0 L-__ -L ____ ~ __ ~ ____ ~ __ ~ ____ ~ __ ~ __ ~

o 5 10 15 20 25 30 35 40

Fractures per metre

Variation of rock mass factor (j) and fracture frequency for chalk (after Hobbs, 1975).

108

Fig. 2.2/16

(a)

l I I \ fIl I I j

I I II .1\/ H \\ I I I I 1/ i ~\'~ 1 1

I I Jr v l\ I T I I \ /) ~ / 1

I I - II I I I \J I / 1 1

1 I '\ ""LA J I r I /'\J IA I T

I - 1 1 (b)

(c)

(d)

Pressure distributions for a circular foundation on a rock masses with diITerent discontinuity set orientations (after Gaziev and Erlikhman, 1971).

109

.-.. E E

o

'-" 1 0 ... C CD E CD

i en

20

o

q q e y

1000

Average bearing pressure (kPa)

Fig. 2~/17 Typical pressure-settlement curve for Grades IV and III chalk (from Burland and Lord, 1970).

110

..... ..... .....

Elevator Tower

~15.750m~

I ct

Silos n -~ , I h I

22.81 Om -.t~22.81 Om----i---22.81 Om-I:>: ct ct· "Co

• ..L...

4 \ • r II~-

Silos 20.0m' Internal Ola.

A

1

Inspection shaft 'C' 1

Silo NO.1

ct

0 I

0.1 0 0 Settlement points

0 0 0 10 0 •

0 0 o ;0 0

Silo NO.2

~ I A 0 .J 0

Silo No.3 Silo No.4

PLAN SHOWING ARRANGEMENT OF SILOS AND INSPECTION SHAFT LOCATIONS

I

610mm Dla. ~ I Columns • \ I 1065mm thick RC silo

: floor slab

" .at 30.0m ~

I :~Extensometer 1-t. borehole

PLAN SHOWING SETTLEMENT LEVELLING POINTS SECTION A-A

Fig. 2.2/18 Silo foundations and positions of levelling points and inspection shafts (after Burland and Bayliss, 1990).

1220mm thick RC raft foundatlor;ls 75mm blinding concrete

o

2

4

6

~8 E --10 .c Q.12 CD

C 14

16

18

20

r-

r-

~

r-

r-

~

r-

-

~

-

-

Fig. 2.2/19

1 Fill VI IV

.... 0_- .. -...... IVN

IV

II

Ill/IV ---.-.... __ . __ ..

IV

Ill/IV

2

VNI

III/IV

III/IV

IV

II

IV

III

Base of rafts

-' 3

VNI

IV

III .-._-_ ... ----_ ..

III

III/IV --•• 0_ •••••• -.-.

IV

IV

. . . . Chalk Grading based on visual inspection

in shafts 1, 2, 3 and 4

4

VNI t---

IV

III/IV ... _---_ .. __ ._.-

III/IV

._------_ .. -----

IV ----------------

IV

Chalk grades based on visual inspection of the chalk beneath the silos in shafts 1, 2, 3 and 4 (after Kee, 1974).

112

0

.-. 10 E E ---.., c CD E 20 CD -=: CD en

30

40

Fig. 2.2/20

Fig. 2.2/21

Load in Silo No. 3 (tonne) 0 2000 4000 6000 8000 10000 12000

I

~

t;:" -~ C") - ~ ,...

t;:" N N ~ -- ~

C") --~ -~ ~ -

0 Av. settlement of silo walls

• Settlement of raft centre ~ -

-~ ,... -~ -~ ~

Settlement of Silo No.3 during first loading (after Burland and Bayliss, 1990).

o

-20 E .s .., c CD 40 E CD E CD

en 60

Silo No.3

" (31/12/73)

,

i ! I

(18}~/74) I

~.~-+~-.----- ... ... " I ..... • -- I ..........

,,,, ! ' .....

.'''''''~(1982) 80 ,

,

4-I~"' ___ ------ 22.Sm

...,

~I

Settlement prorlles for silo No.3 during first loading (after Burland and Bayliss, 1990).

113

..-. E '-"

o o

.c 10 a G> c

20

Fig.2j./22

c o "; "S: "-

Relative settlement (mm) 10 20

Silo No.3 Zero = 100 kPa

• qav = 210 kPa • qav = 250 kPa ... qav = 313 kPa o qav = 353 kPa D qav = 383 kPa

Distribution of settlement with depth beneath the centre of silo No. 3 for various foundation pressures (after Burland and Bayliss, 1990).

Average foundation pressure (kPa) 200 400

8-7

7-6

6-5 '"0 ... r--__ 5 - 4 G> Q.

'*' ,... o o c "iii ... .., o

Fig.2j./23

~ ____ '4~-~3~~--~--__ •

yield bearing pressure

2 -1

Silo No.3

1.80 8 0.42

7 4.81 8.26 6

5 11.34 Depth (m)

14.81 4 3 17.29

19 2 1 23.30

Positions of settlement

measurent points

Vertical strains beneath the centre of silo No.3 (after Burland and Bayliss, 1990).

114

~

~

V1

<t.

Vertical stress (kPa) o 100 200 300 400 500

o I :c;:: x

5

10

-.s 15 .s:: Q. CD C

20

25

30

Fig.2J./24

900- flm . 460

o Bousssinesq

• Soli-structure Interaction analysis

700

600

Relationship between vertical stress and depth beneath centre of silo when filled to capacity (after Nicoletto, 1979).

Soli-structure

I

~o i

i70_! 2~ I . I i

Interaction analysis <t. Bousssinesq

Countours of vertical stress In kPa

Fig.2J./25 Stress distribution beneath a silo when filled to capacity (after Nicoletto, 1979).

...... ..... 0"

18.3m Dla. tank

183kPa 1

10m I

W9 W8

~2~_Y-i WS W6

---C: W7

I Level 1

" , , " I 1 30mOD I W1G

I IV

Level 2 r---

III Twin Marls

Level 3 r-- 20m 00 ---

-I-- -

54 Level 4 51 52 53 II

Approx. water level . ...... _- .......................................... - .. --.- .. -_ ................ ........................ .. ......................... . ........ -.-. . ......... ....................................................................... Y. ..................... c-·····10m·OO············

LevelS Mount Ephraim Marl

II - I

Level 6 Chalk Grade

W1 - 8 Water level gauges 51 - 4 5hafts for vertical deflexlon Instrumentation

Level 1 - 6 Depths at which vertical deflexlons were measured

Fig. 2.2/26 Vertical section showing the position of the instrumentation used to monitor ground movements associated with the tank loading test at Mundford, Norfolk (after Ward et al. 1968).

Om 00

Deflected shape from simple elastic theory

-~-------

Deflected shape of ground surface

Shaft 54

Tat:lk bearing pressure

183kPa

Vertical deflexlon (mm) o

5

.-E

10 -.:.= c '" .. ~ 15 0 -CI) .a 20 .s::. .. a. CI)

C 25

30

Shaft S1

o ._ .. 4--------....

Shaft S2 Shaft S3

Fig. 2.2/27 Section beneath tank showing the distribution with depth of the immediate vertical deflections under maximum tank load at shafts 1, 2, 3 and 4, and the deflected shape of the ground surface (after Ward et aI., 1968).

117

Fig. 2.2/28

.-. E E --c 0 .->< CI) ---CI) -c -co u .-1: CI)

>

.-. E E --c o .;C CI) ---CI) -c "i u t: CI)

>

Bearing pressure (kPa) o 50 100 o ... ~~---r--r--r

0.2

0.4 Level 4

Level 3

0.6

0.8 Level 2

1.0 Shaft S1

Level 1 1.2 Bearing pressure (kPa)

200 100

o~~~ ~ ... -LevelS ------ -0.2 ~vel4

Level 3 / Level 2 Level 1

Shaft S4

Relationship between pressure and immediate deflection at various levels in shafts 1 and 2 (after Ward et aI., 1968).

118

~

E ---.. CD 20 .. m ~ ~

10 0

• .. 0 J:

~ 0.002 ~ ---C 0.004 .-m .. .. t/) - 0.006 m (,) .- 0.008 t: ~

0.010

0.012

Fig. 2.2/29

-0.4

-0.2

0.0 ~

E 0.2 E 0.4 ---.. C 0.6 CD E CD

0.8 -= 0.1 CD en 1.2

1.4

1.6 0

Fig. 2.2/30

200 Time (Hours)

Levels 3-4

Levels 2-3

Short-term. tank test; relationship between time and vertical strain at various levels beneath the centre of the tank during loading, standing and then unloading (after Ward et aI., 1968).

FIlling of the tank complete

S3

Differential between S1 and S2 .

20 40 60 80 100 120

Time (days)

Long-term. tank test; relationship between vertical deflection of level 1 (relative to level 6) in each shaft after completion of loading. The differential settlements at level 1 between the centre (S1) and the edge (S2) of the tank are also shown (after Ward et al., 1968).

119

0

.-. '#I!. e..... 0.01 c --as '-.. rn -as ()

0.02 --t:: ~

0.03

Fig. 2.2/31

Time (days) 0 100 200 300

E = 3980MPa

Level 3-4 (Grade II)

E = 1160MPa Level 2-3 (Grade III)

Level 1-2 (Grade IV-III) E = 488MPa

Long-term tank test; relationship between time and vertical strain for three levels beneath the tank over a period of one year (after Burland, 1975).

120

Fig. 22/32

Visual Description

Lumps less than 3Omm, in soft matrix

Rubbly, friable chalk, joints 10-50mm apart and partly open to 4mm

Rubbly to blocky, joints 3-100mm apart t03mm

Blocky; Joints 30-11 Omm apart, psartly open to 3mm

Blocky, tight Joints III - II

Massive; tight minor joints more than 200mm apart. Some major horizontal joints partially open to 3mm.

~

Chalk Grade

V

IV

IV - III

III

II

Depth below ground level (m)

o

t-- 2

Level of slab test

4

I-- 6

t-- 8

t-- 10

t-- 12

t-- 14

t-- 16

'Profile of chalk grade at Luton, site D (after Powell et aI., 1990).

121

Bearing pressure (kPa)

o 200 400 600 800 1000 o ~~--~------~-----,------~----~~-

2 .-. E E 4 --­.., C CD 6 E CD

i 8 UJ

10

12

Fig. 2.2/33

o

Load maintained

Load maintained

Load maintained for 312hrs

Load maintained for 134hrs

Pressure-settlement curves for the slab loading test carried out at Luton, site D (after Powell et al., 1990).

Bearing pressure (kPa)

200 400 600 800 1000 o ~-------,-------,--------.-------~--------~--

2 .-. E E 4 --­.., i 6 E CD B 8 CD

UJ 10

12

Fig. 2.2/34 Pressure-settlement curve for the slab loading test at Luton, site D with the creep settlement removed (from Powell et aI., 1990).

122

Visual Description

Superficial silty clay

Structureless chalk with Intact lumps

dimensions 3O-6Omm

Soft blocky chalk: vertical and horizontal joints spaced 3O-60mm apart and open from 2-Smm. Several prominant near-vertical joints up to 20mm wide and filled with clay and sand

At 2.3m, 40mm thick sandy clay seam

/ Inclined at 4 degrees to the horizontal

Medium-soft blocky chalk. Vertical joints 6O-150mm apart, open to 2mm

BASE OF PIT

NOTE: Scattered nodular flints and Inclined veins of tabular flint, typical of the Middle Chalk, are found throughout the section

Chalk Grade

V

IV

...................

III

Depth below ground level (m)

o

1

~ 2

...... Level of pad test

~ 3

4

Fig.2j,/35 Profile of cbalk grade at Luton, site C (after Powell et at, 1990).

123

o o

~

~ "'-'" o 0.2 --... ca ... 'E CI) 0.4 E CI)

-= CI) en 0.6

Bearing pressure (kPa) 200 400

Load maintain

~72hrs Load maintained

~ for72hrs

600

~ ::::~\ Load maintained

~ . for72hrs

~~~::!ntalned .... Settlement~

= 10.26mm ~

0.8 -

Fig. 2.2/36

o

~

(ft. "'-'" o 0.2

i ... ... i 0.4 E CI)

-= CI) en 0.6

0.8

Fig. 2.2/37

o

Pressure-settlement curves for the 1710mm pad loading test carried out at Luton, site C (after Powell et al., 1990).

Bearing pressure (kPa)

200 400 600

Pressure-settlement curve for the 1710mm pad loading test at Luton, site C with the creep settlement removed (from Powell et aI., 1990).

124

Fig. 2.2/38

-E E

""-'" ... c: CI)

E CI) -= CI)

en

Fig. 2.2/39

Observations on columns

, With :98m square bases

[!] ~ ~ 'Observations on columns

with 3.35m square bases

/ . '" [!] [!] [!]

10m

Simplified plan showing the arrangement of the 1.98m and 3.35m square footings at Basingstoke (after Lake and Simons, 1975).

Time (Weeks) 0 10 20 30 40

0

-- -----@

10

20 • 3.35m square bases

D 1.98m square bases

Relationship between time and average settlement for the 1.98m and 3.35m square footings (after Lake and Simons, 1975).

125

S1

S22

Fig. 2.2/40

Fig. 2.2/41

• R.C. columns

o Borehole positions

o Trial pit positions

* Levelling positions : •........... Ii 8 BH1 * * ~:. •

r--r-_-c-__ S3 __ S_4~ .·······.L ... j .... J * .............• * ........... * ........... ~ ...... * 1. • . . . ..... I I :

.' : _ _ 1 BLOCK B :

BLO_CK A .0 9 ;'H2 ....... 1.... ............. .. ....... q.: *r· .. * .......... *

10m

* : I : I

• .*: I . I : I

r----~, I I I I I I I I I I I I I IV: I I

I I I I I I . .

Raft joints

Plan of raft foundations showing the position of levelling stations boreholes and trial pits (after Burland et aI., 1975).

0

2

- 4 -E -.r:. 6 c.. CD 8 a -

10

12 -

14

m 1

1 I 1

I 1

1 I

I I I

Fill

Weathered chalk with some flints

Tentative grading Based on SPT N values

V - VI

V

IV - V

Unweathered chalk

Profile of chalk grade based on SPT 'N'values (after Burland et al., 1975).

126

~ ca a. ~ ---G> :; Floor slabs :: and some cladding G> a. 60 C) c: .-'"" ca 40 G> .c G> ~20 '"" G>

~ o 1---4-=

~

E 2 E ---.., c: G> 4 E G> -:= G> 6 en

Backfill and

I Roof and other finishes cladding:

, , I

I I I I

I I I I I I I

Time (weeks) 20 30

•• ' .~ ••. _ ••• ___ H"_~_"' ___ '"

40

S4 S22

Fig. 2.2/42 Relationship between time and settlement for some of the levelling stations in Block A (after Burland et al., 1975).

127

The addition of

thi fai

s block caused

lure

-l-

l

!

28.1t

I

i

~

Centre of gravity of base

I 22t

r i/ Steel bloc ks

!

i 1 ....... - Steel beams 1.1 t

l....

Safety guy

Fig. 2.2/43

'\l ...J

--

W 1..

- ! ....... - 1.ssm Dia . .......

i

! I,

i .~ '-.J -u n

E CJ) Y"'" E . Y"'"

• M Y"'"

N

" []

Cross-section through load test in the base of the cof[er-dam showing concrete cylinder, deep sett1ement points and kentiedge (aller Burland et al., 1983).

128

Load (t) o 40 50 60

o ~~~~----'-----r-----~----~----I

.-.. E E

20

--- 40 .. C G)

E G) E 60 G)

en

80

100

Fig. 2.2/44

2

Load maintained overnight E = 4.4 MPa

no further settlement

o Settlement of point 1.19m below footing

o Settlement of footing No recovery on removal of load

Results of loading test in base of coffer-dam (after Burland et aI., 1983).

1977 1978 1989 1980

Raft completed f.-- Building completed

-.~ ______________ ~ North end

South end 14 _

Fig. 2.2/45 Time-settlement observations on cable chamber (after Burland et aI., 1983).

129

-" w 0

Applied Stress (kPa)

o 100 200

o

~ m11~ Ii!! Ii!! Ii1I ~ -4

0.02

300 o

Applied Stress (kPa) soo

I : o ~~d ~ 3et'11 mf~

Ii!! (fl' . m!ll 1 3 &1ll1

1i1I1 111]1

1000

.-. "'0 IilIIlil 1illD11 .-. 0.1 -fI. 1lI1 fI.

"'-" "'-" f! m3 !ill a: r 1 0 .-~ co a: ~

c: CI)

E CI)

= CI)

en

m1 0 Ii!!~ 1 .-0.04

Ii!! 1 ~ co

rn1 a: ~

c: 1 CI)

1 mi E CI)

0.06 -:= CI)

en

O.OS i "\..\ m6

0.1

(a)

0.2 -

0.3

:., 1 i ~ .:'.: 1 1lIt!J1

4g ll'l1 4rn m 1

1 m 1 1m ElJllI 1

(;31

1mP~ 1 EI..... 1 1 ~ .. 1

EI 1 1m

m1

0.4 -L-_______________ -----'

(b)

1 22.SSm Dla. rafts for silos at Bury St Edmunds, Suffolk 4 1.9Sm & 3.3Sm square footings at Baslngstoke, Hampshire 2 1S.3m Dla. tank loading test at Mundford, Norfolk S 16.Sm & 14.Sm square rafts at Reading, Berkshire 3 3m * 1.Sm slab & 1.71 m Dla loading tests at Luton, Bedfordshire 6 1.55m Dla. In-situ loading test at Salisbury, Wiltshire

Fig. 2.2/46 Relationship between applied stress and settlement-ratio showing pre-yield and post-yield behaviour for full-scale structures and large-scale loading tests.

.-. E E '"-" .. c CD E CD -= CD til CD C) ca ... CD

~

Fig. 2.2/47

Bearing pressure (kPa)

1000 1500 2000 O~~"~~r-------II-------'--------~

5

10

910mm

15

610mm

300mm plate Dia.

20

Pressure-settlement curves for plates of various diameters (after Lake and Simons, 1975).

131

Bearing pressure (kPa) 0 1000 2000 3000 4000

0

Grade II chalk

..-.. 762mm E 5 E --... c CD

i-10 -= CD U)

CD C) cu 15 "-CD

~

20

254mm 152mm Dia. plate

(a)

Bearing pressure (kPa)

0 500 1000 1500 2000 0

Grade V chalk

..-.. E 10 E --... c CD E 20 CD 762mm -= 152mm Dia. plate CD U)

CD C) cu 30 "-CD

~

40 305mm (b)

Fig. 2.2/48 Pressure-settlement curves for plates of various diameters (after Hodges, 1976).

132

-.. as D.. Cl --tJ) ::l -::l "0 0 E .. C as () CD

U)

Fig. 2.2/49

2

1

Loading cycle D:.1 .4 • 2 o 5 • 3

0 0 50 80 120 180 280 400

Plate diameter (mm)

Relationship between modulus and plate diameter of widely jointed Maarstichtian chalk (Nienhuis and Price, 1990).

133

2000

0 Mundford

0 Burland & Lord (1970) ..- 0 Luton as D. q Powell et al. (1990) .:.= ---CD 0 I '- I

:::J aJ 0 '" '" 0 ! 1000 -

Dc:P 0 C. I

C) I

t: I .- 000120 0 '- Mundford Tank as CD m

I

0 Ot§] 0 0

0 o ~ ____ L-__ ~ ____ ~ ____ ~ ____ ~ ____ ~ ____ ~

o

Fig. 2~/50

10 20 30 40 50 60 70

R (%)

Relationship between load-intensity and creep ratio R from plate tests (after Burland and Lord, 1970).

134

10

8

.-.. ~ as

6 "C --E E -CD ..., as ~

Q. 4 CD

CD ~

0

2

o

Burland & Lord

rate ~

1200kPa

1820kPa Bearing pressure

... ... ...

... ...

... ...

... ~

.. ' ............... , ....................... . 10 100 1000

Time (min)

Borehole test (Site D)

(a)

10000

Fig. 2.2/51a Relationship between creep rate and time from plate test at Luton, site D (after Powell. 1990).

135

10

8

.-. ~ m

6 "C --. E E "-" CD ... m ... Q.

4 CD CD ... 0

2

o

Fig. 2.2/S1b

10

Burland & Lord rate

445kPa

100

665kPa Bearing pressure

Trench tests (Site C)

Post yield

1000

Time (min)

(b)

10000

Relationship between creep rate and time from plate test at Luton, site C (after Powell. 1990).

136

2.3 Methods of Assessing the Mass Compressibility of Chalk

Discontinuities have a significant effect on the mechanical properties of rocks

in the mass. The stiffness of discontinuities is generally significantly less than

that of intact rock. It is therefore difficult, if not impossible to make accurate

estimates of mass compressibility from laboratory stiffness measurements

made on intact rock if the rock mass is in any way fractured. Intuitively it

seems likely that the intact stiffness will influence the compressibility

behaviour of the rock mass, but the degree of influence must be related to

the size of the loaded area in relation to the average block size. Hobbs

(1973,1975) suggests that the intact stiffness can be used in conjunction with

the rock mass factor j to obtain rock mass stiffness parameters.

The interrelationships between block size and the dimensions of the loaded

area has been demonstrated by the results of plate loading tests using

different plate diameters in jointed and unjointed chalk (Lake and Simons,

1974, Hodges, 1976 and Nienhuis and Price, 1990). Because the

discontinuities can play an important role in controlling the mass

compressibility it is preferable to perform in-situ tests to assess the mass

compressibility parameters.

The first attempts to understand the load-settlement behaviour of chalk in

the mass was based upon observations of plate tests and foundation

settlements. Wakeling (1966) suggests that the rate of consolidation in intact

chalk has been found to be sufficiently rapid that most structures will

experience full consolidation settlement during construction. It is unlikely that

Terzaghi's theory of consolidation can be applied to chalk since the particles

are cemented to such an extent that the material has a much more rigid

skeleton than soils. Immediate or elastic settlements were thought to

dominate the load-settlement behaviour of chalk and hence settlement

predictions were based on the simplifying assumption that fractured chalk

behaved as a linear elastic material (Wakeling, 1966). However by the late

1960's further observations of the compressibility of chalk under relatively

137

large diameter plates had indicated that yielding occurred at applied stresses

well within the normal stress range applied to full-scale foundations

(Wakeling, 1970, Burland and Lord, 1970).

Mundford represented a milestone in understanding the mass compressibility

of chalk. The work included plate loading tests, a tank loading test, visual

classification and standard penetration tests. The results of the plate and tank

loading tests demonstrated that chalk could indeed meet the stringent

requirements placed upon the settlement of sensitive structures such as a

particle accelerator. It also made engineers aware of the yielding

characteristics of the rock mass. Such behaviour, which does not represent

failure, had not been associated with rock subjected to such low bearing

pressures before. The work at Mundford also highlighted the time-dependent

behaviour of the chalk. Burland and Lord (1970) noted that little attention

was shown to the time-dependent behaviour of chalk.

In comparison to the work done to understand the immediate load-settlement

behaviour of chalk, very little has been done to investigate the time­

dependent settlement characteristics. The longest loading test carried out

remains the Mundford tank test. The tests reported by Powell et al. (1990)

and Powell (1990) are short term in comparison, only lasting a matter of

days.

The mass compressibility of chalk has been investigated using a variety of in­

situ tests ranging from the most popular and cheap techniques such as the

standard penetration test to the less popular and expensive plate loading

tests. Table 2.3/1 shows the in-situ tests commonly used to determine the

parameters defined in Burland and Lord (1970) model (Fig.2.2/18).

138

Table 2.3/1 Methods used to determine the deformation parameters of chalk.

Method

Pressuremeter

Plate loading tests

Geophysics

S.P.T.

Visual assessment

Parameters

./? ?

? ?

139

?

?

Comments

Direction of applied load perpendicular to that applied by a foundation. Results of pressure meter tests carried out in preformed holes (eg Mennard Pressuremeter) will be influenced by distwbance. Self boring pressure meters will have difficulty penetrating flint layers.

Plate loading tests are very expensive to perform and require careful attention to preparation of the plate location and the design of the instrumentation. Plate diameters less than 865mm rarely yield useful data. Interpretation of plate tests results rely on assumptions.

Restricted to seismic techniques (refraction, cross

hole, surface-wave). These techniques are relatively cheap and quick to perform. Surface refraction and borehole techniques will be influenced by anisotropy. Results will be in the form of a modulus depth profile. Some assumptions may be required in the interpretation of raw data.

Tests are cheap to perform. Most commonly

used test to determine ground stiffness profile. Interpretation of test results is based on empirical relationships. Tests results are influenced by drilling method.

Engineering grade classification developed for

Mundford Site is site specific but is used country wide. Relationship between stiffness and grade is empirical. Reasonably large exposures or man­sized boreholes required to enable effective use since correlations with SPT are unreliable.

Pressuremeter Tests

Pressuremeters may be defined as cylindrical devices designed to apply

uniform pressure to the walls of a borehole by means of a flexible membrane.

There are three broad categories of pressuremeter which may be

distinguished in terms of the installation method. These include:

(i)

(ii)

Menard-type pressuremeter (MPM) test - in which the

device is lowered into a borehole (which is usually

slightly oversized).

Self-boring pressuremeter (SBP) test - in which the

device bores its own way into the ground.

(iii) Rock pressuremeter (RP) test - this employs a special

type of self boring pressuremeter that is capable of

boring its own way into rock. Since rock is generally

expected to be much stiffer than soil this pressuremeter

uses instrumentation that has a higher resolution than

the conventional self boring pressuremeter.

Pressuremeters are normally inserted vertically at various depths in the

ground. Figure 2.3/1 shows a typical arrangement for a pressuremeter test.

The test involves expanding a membrane against the surrounding ground by

means of water, gas or oil under pressure. Outward radial deformation of the

ground occurs as the membrane is expands. The object of the test is to obtain

the relationship between the applied pressure and the deformation of the

ground. Ground deformation is determined by measuring the volume of fluid

injected into the central part of the pressuremeter. In devices which are

expanded by gas under pressure (and in some which are expanded by oil) the

radial deformation of the ground is measured directly by a number (usually

3) of 'feeler' arms. In weak rocks the ground deformations are generally small

140

and hence resolution and accuracy of the volume change measurements may

not be adequate for ground stiffness determinations. The use of 'feeler' arms

offer better accuracy and resolution but are still not adequate for stiffness

measurements in weak rocks. The rock pressuremeter makes use of Hall

effect semi-conductors in place of the 'feeler' arms. These permit the

measurement of radial displacements of less than a micron.

Corrections have to be made to measurements of pressure, volume or radial

deformation to account for the effects such as system compressibility,

elevation differences and membrane characteristics. A comprehensive

description of pressuremeters and the principles of pressuremeter testing are

given by Baguelin et al (1978) and Mair and Wood (1987). The nature of the

pressuremeter test is such that within the test zone all soil/rock elements

strained follow very similar stress paths (Jamiolkowski et al, 1985).

Typical data from a MPM test is shown in Fig. 2.3/2. Three distinct phases

are usually evident in tests on soil and are shown in Fig. 2.3/2. These phases

include:

Phase 1

Phase 2

The initial curved portion is attributed to the expansion

of the membrane until it comes into full contact with the

sides of the borehole, also the deformation of any

disturbed or softened zone. In a good quality SBP test in

which the pressuremeter is inserted into the ground with

minimum disturbance, this phase is normally absent.

This approximately linear portion of the curve shown in

Fig. 2.3/2 represents elastic deformation of the

surrounding ground. The onset of phase 2 is considered

to represent the recovery of the in-situ horizontal stress

(Po)'

141

Phase 3 The onset of plastic deformation at Pf in Fig. 2.3/2

marks the beginning of this phase. Phase 3 continues

with the zone of soil in a plastic condition increasing in

radius until a limit pressure Pu is approached. The limit

pressure is the highest pressure that can be sustained by

the ground and can be related to the shear strength of

the ground.

Nearly all of the experience to date in weak rocks has been with Menard

Pressuremeter testing. In certain categories of weak rock, such as chalk, marl

and mudstones the self boring pressuremeter has been used successfully, but

at present experience is relatively limited (Mair and Wood, 1987). Typical

results from MPM tests in moderately weak rock are shown in Fig. 2.3/2. The

shear modulus can be determined from the pressure expansion curve as

follows:

or

The gradient of the linear portion of the pressure expansion curve in phase 2

will give the initial shear modulus (Gi) and the gradient of the linear portion

in phase 3 (if present) will give the yield shear modulus (Gy). In order to

determine Ei and Ey it is necessary to assume an appropriate value of

Poisson's ratio (v) since:

E=2G(l+v)

The yield stress qe is analogous to the creep pressure Pf in Fig. 2.3 /2. The

yield stress qy may be found by extrapolating the linear part of the phase 3

142

pressure expansion curve (if present) back to a vertical line representing the

start of phase 2 in Fig. 2.3/2.

The prime requirement for a MPM test is the formation of a good quality

test pocket with minimum disturbance. Ideally the wall of the test pocket

should be reasonably smooth. In highly fractured chalk it is virtually

impossible to achieve a relatively smooth borehole wall. Irregularities in the

borehole wall will result in limited contact between the membrane and the

wall resulting in non-uniform distribution of stresses and strains. The nature

of the fracturing may cause the pressuremeter to deform asymmetrically due

to movement along joints, or anisotropy, permitting the body of the

instrument (reference frame) to move relative to its original position (Ervin

et al. 1980). Under such conditions the interpretation of pressuremeter data

becomes difficult. Drilling in highly fractured chalk commonly causes

loosening of the ground around the borehole. This type of disturbance can

seriously affect the interpretation of pressuremeter data resulting in Ei being

underestimated. Both the MPM, SBP and the RP test are hampered by the

presence of flint in chalk. Drilling through flints can cause significant

overbreak and disturbance at the level of the flint and below depending on

the drilling method and the care taken by the drill crew. Such overbreak and

disturbance will influence the performance of the MPM test. When rotary

coring in chalk the enormous difference in hardness between chalk and flint

often requires changing the drill bit type to penetrate flints. Since it is not

practical to retract a self-boring pressuremeter in order to change the cutting

tool and since it is unlikely that the cutting tool provided on conventional

SBP's and the RP are capable of cutting through coarse gravel size flints, the

progress of this device will be terminated by the presence of flints.

Some of the difficulties encountered as a result of an uneven borehole wall

may be overcome by performing unload-reload cycles so that Gur can be

determined. Jewell and Fahey (1984) found that for tests in moderately weak

siltstone the ratio Gur/Gi varied between 1.2 and 7 in more weathered

material. The values of Gi were found to vary widely as a result of

143

disturbance whereas Gur values were much more repeatable and consistent.

Hobbs (1970) compared the results of pressuremeter (MPM) tests with that

of plate tests on chalk at Mundford. Hobbs found that the unload-reload

modulus E + values determined in the linear elastic phase of the

pressuremeter test agreed well with the yield modulus Ey values obtained

from plate loading tests, particularly for the more fractured chalk. Hobbs also

claims moderate agreement between Po given by the pressuremeter and qe

from the plate tests. It is likely that Hobbs' results are severely influenced by

disturbance and poor resolution. The pressuremeter used by Hobbs relied

upon the measurement of volume change which is considered to be

inadequate for tests in weak rocks (Mair and Wood, 1987). The agreement

obtained by Hobbs is likely to be coincidence.

Powell et al. (1990) present a comparison of the results of pressuremeter tests

(MPM tests) and plate tests in chalk at Luton (Fig 2.3/3). The pressuremeter

tests and plate loading tests were performed in adjacent boreholes.

Considering the highly fractured nature of the chalk, the agreement between

the limit and ultimate bearing pressures is fairly reasonable. The initial

loading moduli data from the plate tests are fairly close to those back

analyzed from differential settlement measurements of the full scale

foundation, whereas both initial and reload pressuremeter moduli are only a

fraction of the values obtained large-scale plate loading tests. Powell et al.

(1990) attribute these low values obtained from the MPM tests to the

combined effects of disturbance, of the sensitive microfabric of the chalk

during drilling of the test pockets, and the presence of partially open vertical

joints. The chalk between 2 and 3m depth at this site had an unusually open

macrofrabric which was dominated by vertical discontinuities (see Fig. 5

Powell et aI, 1990). Such a structure would be expected to be anisotropic with

Gy > Gh• Since Gh is obtained from the pressuremeter and Gy from the plate

tests the difference in values seen in Fig. 2.3/3 could equally be attributed to

anisotropy.

144

Longworth and Driscoll (1991) through finite element modelling of a 46m

high cutting in chalk found that the field observations of ground

displacements measured during excavation could only be achieved when the

pre-excavation horizontal stresses were assumed to be zero. This does not

seem unreasonable since the chalk at this locality was dominated by open

vertical joints despite the arguments that could be raised concerning the

inability of the finite element method to properly model a discontinuous rock

mass.

Discontinuity patterns in the chalk often make the rock mass anisotropic

(Nunn et aI, 1983, Toynton, 1983). Since discontinuities are an important

factor controlling mass compressibility such anisotropy is likely to give rise to

mass modulus anisotropy. This may be significant in weathered chalk since

the frequency of sub horizontal joints is often greater than that of vertical or

steeply dipping joints. In such a case the initial modulus in the horizontal

direction Eih is likely to be greater than that in the vertical direction E iv•

Theoretically the pressuremeter is the best method for determining the

stiffness parameters for a soil or rock mass since within the test zone all

elements of ground strained will follow similar stress paths. This means that

stiffness parameters can be determined directly from the pressuremeter data

without the need to make use of any assumptions. However, as has been

mentioned in practice there are some disadvantages in using the

pressuremeter, particularly in the chalk. These inc1ude:-

(i)

(ii)

Drilling a hole in preparation for a MPM test will cause

mechanical disturbance of the ground within the test zone and

the resulting stiffness values measured are likely to be lower

than the true values.

Both the conventional self-boring pressuremeter and the rock

pressuremeter will have difficulty in forming suitable test

pockets in chalk containing flint nodules. Indeed it is unlikely

145

that these devises would be able to penetrate the thick tabular

flints found in some horizons within the Upper Chalk.

(iii) In chalk with moderate to widely spaced discontinuities it is

unlikely that a representative volume of the rock mass is tested

using any of the pressuremeters described above.

(iv) The stress applied to the ground from a typical pressuremeter

test is in the wrong direction to adequately simulate that

applied by a foundation. Only in isotropic materials will the

pressuremeter provide meaningful data for the prediction of

foundation settlement. The discontinuity patterns in chalk,

particularly near the ground surface, impart a high degree of

anisotropy to the rock mass.

Plate Load Tests

The plate loading test can be considered the earliest application of in-situ

testing for the evaluation of soil deformability. It is particularly useful in

materials such as fill, boulder clay, hard fissured clays and weak rocks which

cannot be explored by any other in-situ techniques with the exception of

geophysics. Plate loading tests represent the best method for providing mass

compressibility parameters in weak rock. Although expensive such tests afford

the greatest degree of similitude between the test and the field prototype.

The interpretation of the plate loading test however is complicated by the

fact that the rigidity and geometry of the plate results in different elements of

ground beneath the plate following different stress paths. This results in

problems extrapolating from the test scale to that of a full scale foundation.

Plate loading tests may be carried out in shallow pits or in a borehole. The

test involves loading a circular plate (rectangular or square plates are also

used) and monitoring the settlement. Techniques for carrying out these tests

are described by C.P. 2001:1957, Burland and Lord (1970), AS.T.M. Dl194-

146

72, Tomilinson (1975) B.S.5930:1981, Clayton et al. (1982) and B.S.1377:1990

(Part 9). In the test the soil or rock surface to be tested is carefully cleaned

by removing all loose and softened material and the plate is bedded into this

surface using sandi cement mortar or plaster of paris. The cleaning operation

is of particular importance as any loose compressible material left on the test

surface will lead to errors in the determination of initial modulus E j •

Adequate cleaning may be difficult or impossible when tests are carried out

down a borehole.

Load is applied to the plate via a load cell and a hydraulic jack. The jack

may bear against beams supporting kentledge or reaction may be provided by

tension piles or ground anchors installed on either side of the load position.

In weak rocks such as chalk tension piles are necessary if it is intended to

study the post-yield behaviour with an 865mm or larger diameter plate.

When using tension piles it is important that they are positioned outside the

zone of influence of the plate and hence do not influence the results.

B.S. 1377: 1990 recommends that tension piles should not be positioned less

than three times the plate diameter from the centre of the plate. The manner

and rate at which the load is applied to the plate gives rise to two categories

of test:

(i) Constant rate of penetration test (CRP)

The load is applied in a controlled manner such that the selected rate

of penetration is uniform and continuous. The rate of penetration in

CRP tests carried out on the chalk at Luton was 2.5mm per minute

(Powell et al. 1990). At such rates only the undrained or immediate

settlement behaviour is obtained.

(ii) Maintained incremental load test (ML)

The load is applied in a series of discrete increments with a certain

time interval between the application of each increment. Burland and

Lord (1970) suggest that in tests on chalk the time interval between

increments may be based on the rate of settlement. Burland and Lord

147

adopted a settlement rate of 0.005mm per half hour (0.0002mm/min).

The time interval between loading increments was never less than 15

minutes. Clayton et al. (1982) recommend a rate of settlement of

O.004mm per minute. Since the chalk exhibits time-dependant

settlement characteristics the values of Ei and Ey will be influenced by

the time interval between loading increments together with the

magnitude of the increment. The creep rate in chalk is more

significant at stresses above the yield stress and hence the value of Ey

is likely to be influenced more by the interval between loading

increments than Ei . At stresses below the yield stress, Burland and

Lord's rate of settlement is generally achieved within 24 hours (Powell,

1990).

Settlement of the plate is measured using dial gauges or linear displacement

transducers. In order to measure any tilt Clayton et al. (1982) suggest using

four gauges on the perimeter of the largest plate. These instruments are

normally supported on rigid uprights driven (or grouted) firmly into the

ground at a distance of at least twice the plate diameter from the plate

centre. A reference beam for mounting gauges should be adequately

protected from temperature changes. When the test is conducted down a

borehole the vertical settlement of the plate is transferred to the ground

surface by an invar tape reference system described by Ward et al. (1968).

When the test is performed in a shallow pit the settlement of the plate may

be monitored using precise levelling using deep bench marks situated outside

the zone of influence of the plate. Clearly the time taken to perform a round

of levels is such that this technique is unsuitable for CRP tests.

It was noted in section 2.2 that the mass compressibility of fractured chalk is

sensitive to plate diameter. Clayton et al. (1982) suggest that as a general

rule of thumb the plate diameter should never be less than either six times

the maximum soil particle size or six times the maximum intact block size.

B.S.1377:1990 recommends for plate tests on fissured clays that the plate

should be more than five times the average block size. Lake and Simons

148

(1975) suggested the use of a 600mm dia. plate on chalk which had an intact

block size in the range 50-100mm, based on predicted and observed

settlements of a building at Basingstoke, England. Although the plates used

on the chalk in the U.K. range in diameter between 140mm and 1710mm the

preferred size is 865mm. Recently plate loading tests carried out for the

foundations of the new Dartford bridge a 300mm diameter plate was used.

Such a small plate would underestimate the mass compressibility of all but

the most weathered chalk.

Typical results of a plate loading test on chalk are shown in Fig. 2.2/18. The

stiffness of the chalk as determined from the data presented in Fig. 2.2/18 is

given by the equation for a uniformly loaded rigid punch on a semi-infinite

elastic isotropic solid, ie.

where:

E = rcqDp (l-v 2)

4 Pp

E = elastic modulus r

q = applied pressure between the plate and the ground

Dp = plate diameter

v = Poisson's ratio

Pp = settlement under the applied pressure q

When calculating Ey the origin for the applied pressure q is taken at the yield

stress qy (see Fig. 2.2/18).

149

If the plate is considered as flexible then the equation becomes:

or

where: Pay =

Pcentre -

E = O. 844qDp

(1-v2

)

Pav

E = qDp (1-v2

)

Pcentre

average settlement under the applied load q.

settlement at the centre of the plate under the

applied load q.

In the case of a rigid plate P av = P centre' If average settlements are considered

the ratio Eflex:Erigid = 1.08: 1 whereas for settlement measured at the centre

the ratio is 1.27:1. Clearly, since the rigidity of the plate cannot be

guaranteed (particularly with large diameter plates) it is preferable to

determine average settlements from a number of settlement measurements

made at different positions and different distances from the centre of the

plate.

For intact chalk the value of Poisson's ratio is normally between 0.25 and

0.30 at stress levels of 50% of the uniaxial compressive strength (Bell et al.

1990). Bell et al (1990) found that the value of Poisson's ratio increases with

increasing stress level. Table 2.3/2 gives average values of Poisson's ratio at

different stress levels for intact Upper Chalk from Kent. This extends the

overall range to between 0.1 and 004. Clearly the influence of this range of

values on th~ term (1 - v2) will be relatively small and hence will have a

negligible effect on the modulus (Schneider, 1967). Bell et al. (1990) tested

chalks with a relatively wide range of porosities (24 to 47%). Measurements

of Poisson's ratio as part of this study suggest that it is not related to porosity,

150

Table 2.3/2 Values of Poisson's Ratio determined from unconfined

compression tests (from Bell et al., 1990)

% ac Average Poisson's Ratio

7.0 0.170 21.3 0.270 28.4 0.273 42.5 0.293 56.7 0.310 70.9 0.343 81.5 0.367

The above equations assume that the rock mass is an elastic continuum,

whereas in reality it is a dis continuum comprising both elastic and non-elastic

components. This assumption will clearly lead to errors in the determination

of moduli from the plate loading tests when these equations are employed

(Schneider, 1967).

Since E is obtained from load-displacement measurements, an a priori

assumption regarding the constitutive model is necessary. In most cases the

elastic model described above is employed in which homogeneity is assumed.

Chalk, however, often displays a steady increase in modulus with depth

mainly as a result of a decrease in weathering. This non-homogeneity imposes

a serious limitation on the interpretation of plate loading tests if homogeneity

is assumed in conjunction with surface settlement measurements. The effect

of non-homogeneity depends on the ratio Eo/k1 and the diameter of the plate

or foundation. It is not possible from the measurements of plate settlement

alone to determine Eo and k (Hillier, 1992). Since plates of increasing

diameter will stress the ground to greater depths the effect of non­

homogeneity cause the average E determined from plate settlements to be

sensitive to the plate diameter even when it exceeds the average block size by

1 Eo = surface modulus, k = the rate of increase in E with depth

151

some 5 or 6 times. This sensitivity of observed load-settlement behaviour to

plate size has been observed in the results of plate loading tests reported by

Lake and Simons (1975) and Hodges (1976). It appears that Eo and k may

be determined from the results of a suite of plate loading tests in which

plates of various diameters are used. Lake and Simons (1975) presented the

results of plate loading tests carried out with different plate diameters ( 0.3,

0.61 and 0.91m) in chalk at Basingstoke, Hampshire. Hobbs (1975) used these

data to determine Eo and k (Eo = 97 MPa, k = 19 MPa/m). However the

results from the 300mm diameter plate are considered unreliable since the

plate diameter is close to the average joint spacing (50 to 200mm). An

alternative to using plates of different diameter is to use a single plate at a

number of different depths in order to determine a profile of modulus with

depth.

These techniques for determining Eo and k all involve performing a number

of plate loading tests, which will prove expensive. An attractive alternative is

to make measurements of settlement at a number of discrete points below

the centre of the plate (Marsland & Eason, 1973) or on the ground surface

around the plate (Rocha Filho et aI, 1987). An underplate settlement

measurement system has been developed by BRE (Marsland and Eason

1973) to investigate the effects of bedding disturbance and expansion of the

soil on load-settlement behaviour of large diameter plates. The system allows

the settlement at 4 points beneath the plate (set at 150mm in a 38mm

diameter axial borehole) to be recorded via a transducer system housed in

the plate (see Fig. 2.3/4). In this way the settlement distribution beneath the

plate may be measured and the corresponding vertical strains calculated. A

typical set of pressure settlement curves is shown in Fig. 2.3/5.

In order to determine modulus values from underplate strain data, it is

necessary to assume a distribution of total stress beneath the centre-line of

the plate. An axially symmetric stress system may be considered for the

ground between each pair of measuring points. From a suitable starting point

152

(eg the vertical overburden pressure), the (secant) Young's modulus for each

element of ground can be calculated from:

where:

Aav = assumed change in vertical stress

A a h = assumed change in horizontal stress

A €v = change in vertical strain

v = Poisson's ratio

For most chalks it is reasonable to assume v = 0.25, Hence:

However were the vertical and sub-vertical joints in chalk are open it is likely

that Ka approaches zero (Longworth & Driscoll, 1991). This implies that

there are no horizontal stresses transmitted through the rock mass and hence

a change in vertical stress will not bring about a change in lateral stress. In

such a case were A a h = 0, E will be given by:-

This condition is only likely to be true whilst the sub-vertical joints remain

open. The rock mass will therefore behave as a series of columns. As the

vertical stresses are increased, these columns are likely to yield of fail

153

resulting in the development of horizontal stresses. In such a case the above

condition is no longer valid.

At the simplest level the stress changes can be estimated on the basis of

solutions available in the literature for perfectly elastic materials and

idealised geometries. However finite element analysis of plate loading tests

carried out by Hillier (1992) indicates that serious errors in modulus arise if

underplate strain measurements are made within half a plate diameter of the

underside of the plate, (Fig. 2.3/6) when such a simplistic approach is taken

for the determination of the stress changes. It will be seen from Fig. 2.3/6

that for a typical 865mm diameter plate the BRE underplate device will yield

values of modulus that are < 10% in error (at load factors < 0.25) for only

the two deepest elements of ground under consideration. The errors in the

determination of E for the two elements nearest the plate will give rise to

significant errors in the determination of Eo and k. For large diameter plates

a longer underplate settlement measuring device than that currently used by

BRE is required. The use of longer devices may present problems for

installation in chalk particularly when flints are abundant.

It was shown in section 2.2 that the stress distribution beneath the plate will

be influenced by both the orientation and persistence of the joints. Model

tests on jointed rock masses all show that the stress distribution beneath the

loaded area is significantly different to that based on an elastic continuum. In

a real rock mass such as the chalk the situation is further complicated by the

fact the style of jointing (orientation, persistence and aperture) often changes

with depth. It is arguable, therefore, that the stress at a given point below the

plate cannot be predicted with any accuracy, and this casts doubt on the

validity of modulus determinations from sub-plate deformation

measurements. These arguments concerning stress distributions in jointed

rock will apply equally to the interpretation of pressure meter data and

instrumented foundations.

154

Rocho Filho et al. (1987) proposed a method of determining Eo and k from

the form of the distribution of vertical surface settlement outside the area

loaded by a circular rigid plate. The form of the surface settlement depends

upon the degree of inhomogeneity and Poisson's ratio. Thus when Poisson's

ratio is known the ratio of Po (the settlement at a given radius outside the

plate) to Po (the settlement inside the plate) will lead directly to a value of

Eo/kDp, where Dp is the diameter of the plate. Eo can be determined from

the magnitude of the plate settlement for a given loading increment.

Rocho Filho et al.'s method is based on elastic theory and uses the numerical

model developed by Brown and Gibson (1972). This model assumes an

elastic continuum. The rock mass, however, is not a continuum, and the form

of the distribution of surface settlement will depend to a large extent on the

geometry of t?e discontinuities (Hungr and Coates, 1978).

The variation in modulus with depth may be determined from a series of

plate tests carried out in a borehole at different depths. However the method

has the following limitations:

(i)

(ii)

The maximum depth will be limited by the position of the water

table since the base of the hole requires minimal disturbance

and careful preparation.

The maximum size of plate will be limited and hence the test

data may not be representative of the rock mass.

(iii) The accuracy of the modulus-depth profile will be limited if the

modulus increases rapidly with the zone of influence of the

plate.

Accurate settlement predictions of full scale structures based on plate loading

test data requires knowledge of the modulus depth profile. The average E

derived from surface measurements of plate settlement alone can give rise to

155

significant errors in settlement predictions when the ground is non­

homogeneous. If the plate is the same size as the foundation or the ground

has a uniform stiffness such errors do not arise. In making settlement

predictions it is sometimes assumed that:-

Where

Pc = settlement of foundation

Pp = settlement of plate

Be = width/diameter of foundation

Bp = width/diameter of plate

a = empirical factor

The empirical factor a may be determined from tests with different size

plates. However Hobbs (1975) shows that except at values of a approaching 1

for relatively small foundations the above equation does not adequately

model the condition of steadily increasing modulus with depth, which is the

only condition apart from homogeneity in which this procedure can be

realistically applied.

Plate loading tests may be used to measure the creep behaviour of the rock

mass. However these tests by there very nature are time consuming and

hence expensive. When considering the measurement of creep settlement the

accuracy and stability of the displacement measuring system becomes

crucially important. Large diurnal temperature variations (which can be as

much as 20°C in the UK) combined with high winds cause great difficulty in

achieving stability, particularly using electronic instrumentation. In order to

achieve reliable measurements of creep over long periods of time it is

necessary to thermally isolate the measurement system or devise a suitable

method of compensating for the effects of changes in temperature. The use of

precise optical levelling may overcome the temperature effects to a large

extent but suffers from stability problems in high winds and a lack of

156

resolution. Resolution is a particular problem in attempting to accurately

measure settlements at stress levels below yield. The transducer systems used

by BRE and described by Ward et al. (1968) and Burland and Lord (1970)

are capable of providing sufficient resolution but suffer from temperature

instability. This can produce significant errors when the settlements measured

are very small. The result is that it is very difficult to obtain reliable time

dependent settlement data at low bearing pressures (ie <300kPa).

Geophysics

Geophysics provides a number of indirect methods of determining ground

stiffness. Such indirect methods do not involve the direct measurement of

stress and strain but make use of mathematical relationships to determine

stiffness parameters such as E and G. Geophysics makes use of variations in

the physical properties of the ground. Since the propagation velocity of

seismic waves through the ground is related to density and elastic constants

seismic techniques are most appropriate for determining stiffness.

Geophysical methods such as the measurement of longitudinal and shear

wave velocities by down-hole and cross-hole techniques are of relevance to

the measurement stiffness in soils and weak rocks (Miller et al. 1975, Bodare

and Massarsch, 1982). These measurements also provide an insight into the

stress-strain behaviour the ground at low strains (y ~ 10-5%). The design and

evaluation of structures for earthquake excitation has prompted increased use

of seismic field studies to determine in-situ moduli at low strain levels. The

stiffness of rocks and soils determined from seismic methods are frequently

significantly greater than those determined from monotonic loading tests in

the laboratory. Recent research however has shown that intact soil

specimens are much stiffer at small strains (€ = 0.001%) than at larger

strains measured with conventional laboratory instrumentation (ie € = 0.1 -

10%) (Jardine et aI., 1984, Jardine et al. 1985). Fig 2.3/7 shows a typical

relationship between secant modulus and strain. The maximum strains which

have been observed in the ground adjacent to full scale structures are often

157

at the bottom end of the range of resolution for conventional laboratory

instrumentation and in many cases the working strains are at least an order

of magnitude less than that which can be measured by conventional means

(see Fig. 2.3/7).

Due to the small strains associated with the propagation of seismic waves

only the initial modulus (Ei) for chalk may be determined with geophysics.

The fact that seismic techniques are unable to give any indication of the yield

stress or yield modulus may be considered a serious disadvantage. However

this should be set against the fact that a profile of Ei with depth can be

provided relatively rapidly, and for a representative volume of ground.

Moreover some seismic methods do not necessitate the use of boreholes.

The small-strain moduli are estimated from determinations of P and S wave

velocities taken together with the in-situ assessed bulk density and Poisson's

ratio. Table 2.3/3 lists the principal types of elastic wave and how the

propagation velocity is related to elastic modulus.

It will be seen from Table 2.3/3 that the determination of Young's modulus,

E, from the P wave velocity, Vp' requires Poisson's ratio to known for the

rock/ soil mass. Since this is a difficult parameter to measure in-situ it is often

estimated. The determination of E becomes impossible when v = 0.5, a

typical condition in near-surface saturated ground. It is therefore more

attractive to measure the S wave velocity V S' since the relationship between

shear modulus G and Vs does not involve Poisson's ratio and G is the same

for both drained and undrained loading (Grainger et al., 1973).

There a number of different seismic techniques that are used to obtain a

modulus depth profile in soils and rocks. These methods include:

158

-" V1 ~

Table 2.3/3 Principal types of elastic wave

Wave Designation Remarks Propagation velocity

Compressional, dilational P Particle motion in direction of propagation. Vp2 = E(l - v)/p(1 - 2v)(1 + v)

(infinite medium)

Distortional, shear S Particle motion normal to direction of propagation. V/ = G/p

Vertically polarized SH

Horizontally polarized SV

Rayleigh wave R Retrograde elliptical motion at surface. Vr = C*Vs

(C = constant which depends upon Poisson's ratio v)

Love wave L Particle motion normal to direction of propagation in Short A: Vt = (G/PtY'~ plane of interface. Long A: V2 = (Gip2)~

Stonely wave (generalised Rayleigh Surface wave in elastic half-space where two layers O.988Vs wave) have similar shear wave velocities.

V p = propagation velocity of P waves, Vs = propagation velocity of S waves, Vr = propagation velocity of Rayleigh waves, E = Young's modulus, G = shear modulus, v = Poisson's ratio, p = bulk density, A = wavelength, subscripts 1 & 2 refer to layer numbers.

(i) Seismic Refraction

(Grainger et aI., 1973, Abbiss, 1979)

The conventional surface refraction survey is performed using an

impact or explosive energy source in conjunction with a multiple array

of geophones (Fig. 2.3/8). Transit times between source and detector

are provided by a suitable seismograph. These are interpreted by

plotting the transit time against the distance from the energy source

for each geophone. Both compressional and shear waves may be used

to determine the P and S wave velocities respectively. P waves only

have limited use in soils and in such cases S waves are used. The

elastic moduli are determined from the seismic velocities. Typically the

seismic refraction method will allow the ground to be divided up into

layers each having a characteristic P and S wave velocity. Unless the

survey is designed carefully the resolution in terms of a modulus depth

profile will be coarse giving average modulus values for relatively thick

layers. The method will only detect layers when the velocity increases

with depth. In the case of a highly compressible material underlying a

much stiffer material the compressible material will not be detected

since its velocity will be less than that of the overlying layer.

(ii) Crosshole

(Thompson et aI. 1993)

Crosshole surveys are carried out using two or more boreholes (Fig.

2.3 /8). The boreholes are cased with plastic casing which is grouted in

place to provide good acoustic coupling between the borehole and the

ground. A seismic source is placed in one of the boreholes and a

receiver in another at a known elevation (see Fig. 2.3/8). The transit

time for seismic waves to travel between source and receiver is

measured with a suitable seismograph. By measuring the transit times

at different elevations a modulus depth profile can be built up. Both P

and S waves may be used for crosshole surveys although S wave

surveys are more common. P waves are generated using a sparker and

160

are detected using hydrophones in water filled-holes. S waves are

generated using a hammer clamped against the casing and are

detected with clamped three component geophones. The conventional

S-wave hammers produce vertically polarized shear waves together

with compressional waves. In order to determine accurately the transit

times it is necessary to trigger in seismograph and the energy source

simultaneously. This is not always possible particularly with

conventional S wave hammer sources. In such circumstances two

receiver boreholes are required such that transit times can be

measured between receivers instead of between source and receiver.

(iii) Uphole/Downhole

(Sigismond et al. 1983)

In this type of survey the transit times of P or S waves are travelling

between points in a borehole and a point on the ground surface near

the top of the hole are measured (see Fig. 2.3/8). The terms uphole

and downhole refer to the general direction in which the seismic waves

are travelling (eg uphole refers to an energy source located in the

borehole and the receiver located on the ground surface). Velocities

may be determined from a plot of depth against transit time. The

downhole method is used in preference to the uphole method since a

more powerful energy source can be used (without the possibility of

damaging the borehole wall) and the seismic noise is less at a

downhole detector. The seismic waves used in this type of survey

travel in a near vertical direction.

(iv) Surface-Waves

(Abbiss, 1981)

Surface-waves (Rayleigh waves) are generated by a controlled

vibratory energy source. The surface waves are detected by a series of

geophones placed in line with the vibrator (Fig. 2.3/8). The time-

161

domain data collected by the geophones is converted to the frequency

domain using a Fast Fourier Transform. The frequency domain data is

used to calculate the phase velocity and the wavelength of the surface

wave. By collecting data over a range of frequencies (typically 5 to

200Hz) it is possible to build up a velocity depth profile. If Poisson's

ratio is known it is possible to determine the shear wave velocity from

a Rayleigh wave velocity. Thus a shear wave velocity-depth profile can

be obtained which can be easily converted to a shear modulus-depth

profile. A comprehensive description of the field techniques, data

processing and interpretation for continuous surface-wave surveys is

given in Chapter 4.

In most cases the seismic velocities are determined from transit times

measured between the energy source and receivers and the distance travelled

by the wave between source and receiver (not always a straight line). In the

case of conventional seismic refraction, cross-hole and down-hole surveys, the

source imparts a pulse of energy into the ground and hence the pulse velocity

is determined. In the case of surface-wave surveys a continuous vibratory

source is employed. The seismic velocity in this case is determined from the

phase difference between neighbouring receivers. The velocity determined is

a phase velocity which may not have the same magnitude as the equivalent

pulse velocity.

Seismic refraction was used in an attempt to correlate degree of fracturing

with P-wave velocity at Mundford, Norfolk by Grainger et ale (1973). Since

the Mundford site had been subjected to a comprehensive sub-surface

investigation which included the visual inspection of the chalk at different

depths as well as in-situ stiffness determinations (from plate loading tests)

there was ample ground truth to correlate with seismic measurements. A total

of 10 seismic refraction lines were employed over the whole site with 5

located in the vicinity of the tank loading test. The interpretation of the

seismic data together with ground truth data are shown in Fig. 2.3/9. It will

be seen that the seismic refraction survey clearly shows that the velocity

162

increases with depth in well defined steps, which broadly correspond to the

classification system adopted by Ward et al. (1968) which is based primarily

on fracture spacing and aperture. This is generally the case for the other 9

seismic lines. It should be pointed out that since the refraction survey was

designed to give a depth of penetration of 30m it was impossible to

adequately resolve the drift deposits from the highly disturbed grade V chalk.

Hence the correlation with ground truth in these upper layers is not good

(see Fig 2.3/9).

It seems entirely reasonable that the P wave velocity should be influenced by

the degree of fracturing in the rock mass since the acoustic coupling across

fractures will be poor (Wyllie et al. 1956) thus reducing the seismic velocity.

This effect, however, is markedly reduced by the introduction of water into

the fractures. The P wave velocity in intact porous rocks is almost constant

over a wide range of saturations (about 10-95%). Grainger et al. (1973) found

that the seismic interpretation was unaffected by the position of the water

table when the fracture spacing was greater than 60mm. However in the

structureless grade V chalk they found that the velocity was raised from

700m/s to 195Om/s, by saturation. This material tends to comprise blocks of

intact chalk in a groundmass of remoulded (putty) chalk. The skeleton of this ,

material is highly compressible and hence behaves like a soil. When a soil of

low permeability is loaded it experiences no change in effective stress in the

short term if confined. A compressional wave propagating through a soil will

subject the soil to a similar undrained loading even in soils of relatively high

permeability. Since the applied loading will be carried by the excess pore

water pressure and not by the soil skeleton the seismic wave will tend to

propagate through the pore water. The velocity of such a P wave will be

strongly related to the velocity of water (150Om/s) and not to the density and

stiffness of the soil skeleton. This is a serious limitation to the use of P waves

it compressible porous media such as soils. It is because of this limitation that

S waves are now used extensively for the seismic investigation of such

materials.

163

The seismic studies carried out at the tank site comprised a P wave refraction

survey. Abbiss (1979) carried out seismic refraction survey at the tank site

specifically to determine the variation of stiffness with depth. A typical time

distance graph from this survey is shown in Fig. 2.3/10. The time distance

graph shows a continuous curve indicating a continuous increase in P wave

velocity with depth. Such data can be interpreted by fitting an inverse sinh

function to it which corresponds to a linear increase in velocity with depth

(Dobrin, 1960). Such an interpretation has yielded the dynamic moduli shown

in Fig. 2.3/11. It will be seen from Fig 2.3/11 that there is a linear

relationship between the dynamic moduli derived from the refraction survey

and the short-term static moduli deduced from the tank loading test such that

Estatic = 0.4 7Edynamic.

The time increment for each stage of the tank test was several orders of

magnitude greater than the time for each strain excursion associated with the

propagation of a compressional wave through the rock mass. Since the chalk

is known to display time dependent characteristics it seems reasonable that

the deformations measured in the tank test will be more influenced by time

than those associated with the seismic refraction survey and the latter would

be expected to yield the higher modulus. This assumes that the modulus is

independent of strain. Since it generally accepted that the modulus reduces

with increasing strain, this difference in strain level between the static and

dynamic tests could contribute to the difference in moduli. The difference in

frequency between the static and dynamic tests could also influence the

elastic moduli derived from each (Abbiss, 1979). The dynamic moduli

obtained from the seismic refraction survey' when converted to equivalent

static values using the relationship obtained from Fig 2.3/12 show good

agreement with the moduli derived from the tank test. This suggests that the

dynamic stiffness values derived from seismic velocity measurements must be

reduced by an appropriate factor in order to provide equivalent static values

that will not significantly underpredict foundation settlements.

164

Borehole seismic methods have been used to obtain modulus-depth profiles

in many rock types including weak mudstones (Thompson et al. 1993) and

chalk (Sigismond et al. 1983).

Sigismond et al. (1983) present the results of cross-hole and down-hole tests

in chalk which were performed for the site investigation for a power station

in Nogent-sur-Seine, France. Fig. 2.3/13 shows the shear modulus depth

profiles derived from of cross-hole and down-hole shear wave surveys. It will

be seen that the moduli determined from the different surveys show

reasonable agreement below 20m depth. This suggests that the rock mass

does not display a high degree of anisotropy. Above 20m in the putty chalk

and overlying alluvium the two types of survey yield distinctly different

moduli values for a given depth. This may be due to anisotropy. The authors

reported problems in interpreting the cross-hole data within the putty chalk

(10-16m) due to refraction in the overlying and underlying layers which both

have a higher velocity. ,

Below 20m in Fig. 2.3/13 the shear modulus is seen to increase with depth

from 2000MPa at 20m to 3000MPa at 80m. Such an increase in modulus in

chalk would normally be attributed to a reduction in the frequency and

aperture of discontinuities with depth.

Surface-wave techniques offer an attractive alternative to the above method

for obtaining a modulus-depth profile. The main attraction is that a profile of

shear modulus with depth may be determined rapidly without the need for

boreholes.

Geophysics measurements are made at shear strains of about 10-5%. To

obtain assessments of shear moduli at higher strain levels a strain-dependent

moduli model needs to be used. Of the several models that exist that by

Hardin and Drnevich (1972) has been found to be the easiest to use. Powell

and Butcher (1991) used this model to compare results of direct

(pressuremeter, plate, triaxial) tests with indirect (refraction, surface-wave

165

and laboratory piezobender) for a number of different soil types including

glacial till, heavily overconsolidated clay and soft clay.

The Standard Penetration Test

The standard penetration test is perhaps the most common in-situ test used

in the UK It is common practice to include standard penetration tests in

almost all subsurface investigations in soil and weak rock. The test involves

measuring the number of blows necessary to drive a standard penetrometer

through a given distance (300mm) into the base of a borehole using a

standard mass falling through a given distance. Since there is likely to be

disturbed material in the base of the borehole it is standard practice to drive

the penetrometer 150mm before the main drive described above commences.

The number of blows required to drive the penetrometer the standard

distance in the main drive (30Omm) is referred to as the SPT 'N' value. The

test is described in detail in BS 1377:1975 & 1990.

The SPT may be carried out in accordance with BS1377:1975 using a

standard split spoon sampler tool in most soils or a solid 600 cone for use in

gravels or gravelly sand. However BS5930: 1981 extends the use of both the

split spoon and solid cone to rock despite the fact that contractors have been

carrying out standard penetration tests with both types of tool for many years

prior to 1981. Where flints are likely to be encountered in the chalk,

contractors are likely to use the solid cone since flints can cause considerable

damage to the shoe of the split spoon sampler. Neither British Standard

requires the contractor to state the tool used in the test. It is likely that the

SPT N value is effected by the tool used. Montague (1990) compared the

results of tests performed in chalk with both types of tool. Variations of the

order of 2:1 were identified between Nsolid cone and Nsplit spoon at common

depths at the same site. Little work appears to have been done to ~vestigate

the influence of the type of tool on the N value.

166

The popularity of this tests has led to a wide variety of empirical methods

which permit the prediction of strength and stiffness for soils and weak rocks.

The nature of the tests lends itself more to the measurement of strength than

that of stiffness. A number of correlations between 'Elastic Modulus' and SPT

'N' value have been proposed for the chalk over the past 30 years. Most are

based on moduli determined from plate loading tests and very few on the

performance of full scale structures (Wakeling, 1966, Wakeling, 1970, Lake

and Simons, 1970, Kee and Clapham, 1971, Powell et al. 1990). All should be

treated with caution.

It seems reasonable to assume that the penetration resistance of intact chalk

may be related to the intact porosity. Since the mechanical properties of

intact chalk are controlled largely by the intact porosity a good correlation

may exist between the SPT 'N' value and the intact stiffness of chalk. If the

intact strength (which is related to intact stiffness) of chalk influences the

compressibility of the rock mass, then the SPT 'N' value may reflect mass

compressibility where the joint frequency is sufficiently small.

Wakeling (1966) produced a tentative correlation between SPT 'N' value and

'Elastic Modulus' for soft chalk (Fig. 2.3/14). This correlation is based on

only three values of modulus obtained from field tests:-

a pile test (Newbury),

a large foundation (Medway Bridge)

and a plate loading test (Norwich).

Such a correlation is indeed tentative when the background to each case is

considered (see Table 2.3/4). Although Wake ling was careful to take account

of the depth of the foundation in the determination of the modulus he was

unaware of the influence of the size of the loaded area in relation to joint

frequency, on modulus. Tomlinson (1966) raised this point in the discussion

of Wakeling's paper. In the case of the plate loading tests carried out at

Norwich the plate was too small to yield any meaningful results. In the case

167

.... O'l 00

Table 2.3/4 Basis of Wakeling's tentative correlation between SPT 'N' and elastic modulus (from Wakeling 1966)

Site

Newbury

Medway Bridge

Norwich

Foundation

Underreamed pile. Shaft Dia. 1372mm, underream Dia. 2591mm.

Bridge pier foundation. 9.5m • 32.3m

Deep plate-bearing test. Plate Dia. 140mm. 23 tests carried out at 1.5m intervals to a depth of 12m in 4 boreholes.

Equivalent modulus (MPa)

138 at 300 tonnes. Calculated assuming all the load was taken on the base of the pile.

207

117 (average from 23 plate tests)

SPT details

SPT 'N' values from two boreholes increased with depth, particularly below the base of the pile .

No Standard Penetration Tests carried out. SPT 'N' values were estimated by comparing the moisture contents at this site with those prevailing at Newbury and Norwich.

Tests carried out to a depth of 12m in all 4 boreholes. Chalk displayed a uniform penetration resistance with depth.

of the Medway Bridge no SPT 'N' values were measured. SPT 'N' values

were interpolated from those measured at Norwich and Newbury on the basis

of moisture content and obviously the validity of this approach is

questionable (Rodin, 1966). With regard to the pile test at Newbury since

shaft resistance was ignored it is likely that the modulus has been

overestimated.

Wakeling carried out standard penetration tests at Mundford and was able to

improve the correlation discussed above by making use of the moduli derived

from the plate and tank loading tests described by Ward et al. (1968). The

SPT data used by Wakeling came from 4 boreholes, advanced by light

percussion boring, located adjacent to the tank loading test and plate loading

tests T3 and T4. The correlation between SPT N value and modulus is shown

in Fig. 2.3/15. The values of modulus were interpolated since the level at

which they were measured did not correspond to the level of the SPT results.

The two lines A and B shown in Fig. 2.3/15 relate to values of Ey and Ee

respectively. It will be seen from Fig. 2.3/15 that the data used by Wake ling

to support line A is influenced by the incorporation of data from plate tests

in small-diameter boreholes, which are known to be unrepresentative in chalk

due to scale effects and base disturbance.

Lake and Simons (1970), inspired by the tentative correlation between SPT N

value and modulus proposed by Wakeling (1966), developed another

correlation based on 140mm diameter plate loading tests in chalk on the M4

at Welford Theale, Berkshire. This correlation is shown in Fig. 2.3/16

together with that of Wakeling (1970). The plate loading tests were

conducted at depths between 4 and 20m in boreholes advanced by light

percussion boring methods. In order to minimise disturbance a conventional

auger was employed to advance the hole when within 305mm (1ft) of the test

level. At the test level a specially made flat bottom auger was used to trim

and clean the base of the hole. The method used may reduce mechanical

disturbance of the chalk at the test level, but in such small diameter holes it

is impossible to clean the base of the hole adequately and hence the results

169

of these tests will be affected by base disturbance. The plate loading tests

were of the constant rate of penetration type and the rate (1 to 2 nun/min)

was such that the initial modulus Ei could not be determined.

Kee (1968) initially re-analyzed the data used by Wake ling (1966) and

suggested a correspondingly lower correlation between the modulus and SPT

N value. Kee and Clapham (1971) modified this relationship (see Fig. 2.3/16)

on the basis of a number of 200nun diameter plate tests carried out in

boreholes, and the analysis of data from two case histories, Tomlinson (1966)

and Palmer (1966). Tomlinson (1966) gave the results of 143nun diameter

continuous rate of penetration plate loading tests for Addenbroke's Hospital,

Cambridge and Palmer (1966) discussed the results of a 445mm diameter

plate loading test at Purfleet, Essex.

The correlations proposed by Lake and Simons (1970) and Kee and Clapham

(1971) should be treated with caution since both are based on small diameter

plate loading tests. The use of small diameter plates (in general < 865mm)

are known to be unrepresentative in chalk for reasons outlined earlier. Powell

et al. (1990) proposed a correlation based on 865mm plate loading tests and

the back analysis of a full-scale foundation at Luton, Bedfordshire. This r

correlation represents a lower bound of moduli based on settlements of 0.5 %

of the plate diameter and gives modulus values which are generally higher

than Lake and Simons and Kee and Clapham.

Ward (1970) emphasised the fact that the SPT values for all varieties of chalk

seem to vary by only 3 or 4 times and correspond to a variation in E values

of some 100 times. Ward concluded that the SPT cannot therefore provided a

sensitive index of the elastic modulus. This results from plotting the moduli

on a logarithmic scale. Wakeling (1966) introduced the logarithmic scale in

the first correlation between modulus and SPT N value in order to display

the wide range of moduli values. The logarithmic scale was again adopted

Wake ling (1970) for his interpretation of the data from Mundford. Certainly

the Mundford data would appear to justify the use of a logarithmic scale for

170

Young's modulus. This led other workers (Lake and Simons, 1970, Kee and

Clapham, 1971 and Powell et al. 1990) to adopt the same approach even

though the range of modulus values is much less than those observed by

Wakeling.

Stroud (1988), recognising that both stiffness E and N vary with mean

effective stress for soil, proposed that the ratio E/N60 should be considered

together with its variation with strain level or degree of loading by the ratio

q/qult" The parameter N60 used by Stroud is the SPT N value corrected for

60% of the free fall hammer energy (Clayton, 1990b). The variation of E/N60

with 'bet/q(ult)u for the Chalk is summarised in Fig. 2.3/17 based on available

data from in-situ loading tests of shallow foundations, piles and large plates.

'bet represents the average net effective bearing pressure. Values of q(ult)u for

weak rocks and chalk were determined using bearing capacity factors and the

undrained shear strength of the rock mass. The undrained strength may be

estimated from N60 values and the appropriate value of f1 using the following

relationship:

Stroud suggests that for chalk f1 = 25kPa.

Stroud suggests that a value of E/N60 of 15 MPa is broadly consistent with

data presented by Hobbs and Healy (1979) for the base performance of 30

pile loading tests and a number of large scale footings with degrees of

loading up to 0.15. He indicates that a conservative estimate of stiffness for

chalk under a moderate degree of loading (> 0.15) is given by E/N60 of

about 5.5 MPa. Both correlations are shown in Fig. 2.3/16. The choice of

which relationship to use depends upon whether pre-yield or post-yield

settlements are required. It is therefore necessary to know what the value of

the yield stress qy. This cannot be predicted from the Standard Penetration

Test.

171

The relationships between SPT N value and Young's modulus is unusual

since it is linear. However it will be seen from Fig. 2.3/16 that these

empirical correlations provide a much better agreement with Powell et al.'s

data than the logarithmic relationships discussed earlier. This suggests that

Stroud's correlations may give more accurate predictions of foundation

settlement than the other commonly used methods such as Wakeling (1970)

and Kee and Clapham (1971).

The reliability of the SPT in providing accurate stiffness values is made even

worse by the fact that the results are affected by the presence of water, fallen

debris at the base of the borehole, the amount of mechanical disturbance

below the base of the hole caused by drilling and the type of tool used. The

mass compressibility of chalk is a function of the fracture state (orientation,

frequency and aperture) and the mechanical properties of the rock material.

In order for there to be a unique correlation between modulus and SPT N

value these factors must be reflected in the penetration resistance. In highly

fractured chalk a large proportion of the test volume may be through open

discontinuities, whereas in chalk were the discontinuities are closed and over

200mm apart the intact rock will dominate the penetration resistance.

Clayton (1978) demonstrated that the SPT N value is sensitive to the intact

dry density of the chalk. It is likely, however, that in a closely jointed chalk

rock mass the joints would have a significant influence on the SPT N value. A

situation could arise in which a chalk with a high dry density and closely

spaced joints gives a lower N value than a chalk of low dry density and widely

spaced joints. In this context it is the vertical joint spacing and aperture which

are of importance. When the vertical joints are closely spaced and open there

is likely to be less resistance to penetration since the rock is able to deform

laterally with ease ahead and around the SPT tool. The low N values derived

from such a rock mass could lead to a significant overestimation of the mass

compressibility. Where both the vertical and horizontal joints are widely

spaced the SPT N value is likely to reflect better the intact dry density of the

rock and the rock mass compressibility.

172

Evidence of drilling method influencing SPT N values in chalk is given by

Mallard (1977). Fig.2.3/18 shows the results of standard penetration tests

performed by four contractors at the Liitlebrook D power station site.

Contractors A, Band C used the conventional light percussion method of

boring whereas contractor D performed tests in BX rotary percussive

boreholes. It will be seen from Fig. 2.7/34 that contractor D shows a fairly

tight band increasing in N value with depth up to about 6m then a broad

band of essentially constant N value with depth. These trends cannot be

identified in the other contractors results due to the large amount of scatter.

A comparison between contractor C and D indicates that the N values

derived from light percussion boreholes are lower than those derived from

the rotary percussion holes. This possibly reflects the relative amounts of

disturbance ahead of the bottom of the hole produced by the different

drilling methods. The evidence is by no means conclusive but it does suggest

that the SPT N value may be sensitive to drilling method. There have been

no systematic investigations of the influence of drilling method reported in

the literature.

Although the SPT is the most popular in-situ test for providing mass

compressibility parameters for the chalk, it is perhaps the least reliable. The

empirical nature of the correlations between N value and E together with the

fact that the test is insensitive to relatively large variations in stiffness should

cause the results to be treated with caution. The fact that the results of SPT

tests appear to be affected by drilling method and the type of tool used in the

test means that the data can only be usefully employed for profiling rather

than for the determination of design parameters.

VlSUlll Assessment

The systematic collection of geological data and its correlation with

engineering experience and measurement can lead to improved methods of

predicting the behaviour of rock masses. Thus Moye (1955) developed a

method of correlating geological occurrence and behaviour for rock in the

173

Snowy Mountains Scheme. The major part of the Snowy Mountains Scheme

was the construction of tunnels in granite, and the principal geological feature

which controlled the behaviour of the rock mass was alteration of the granite

by chemical weathering. Moye (1955) devised an engineering grade

classification system for the granite in order to ensure that geologists

descriptions were meaningful in engineering terms. It was important that the

various degrees of chemical weathering were recognised and hence Moye's

classification is essentially a classification of weathering of rock material in

engineering terms. This classification system was the forerunner for numerous

other weathering classifications dealing with both the rock material and the

rock mass (eg Ruxton and Berry, 1957, Anon, 1977). Indeed the interest in

weathering classifications has been such that true engineering grade

classifications become insignificant in comparison and some are referred to as

weathering classifications simply because weathering happens to be an

important feature controlling the behaviour of that particular rock mass.

In general engineering grade classification schemes rely on qualitative or

semi-quantitative observations of the rock mass. They are usually developed

for a particular project in order to provide some uniformity in rock mass

descriptions as well as allowing engineering parameters measured at discrete

points or zones to be extrapolated to other areas of the site. Such

classification schemes are thus site specific and should not be confused with

rock mass classification schemes which serve to quantify experience so that it

may be extrapolated from one site to another. These classification schemes

seek to assign numerical values to those properties or features of the rock

mass considered likely to influence its behaviour, and to combine these

individual values into one overall classification rating for the rock mass.

Rating values for rock masses associated with a number of mining and civil

engineering projects are then determined and correlated with rock mass

behaviour. Goodman (1976) and Hoek and Brown (1980) have reviewed a

number of rock mass classification schemes that have been developed for a

variety of purposes, although the majority are for tunnelling applications. Two

of these schemes, the NGI tunnelling quality index (Q) developed by Barton

174

et al (1974) and the CSIR geomechanics or Rock Mass Rating (RMR)

scheme developed by Bieniawski (1973, 1976), are currently widely used in

civil engineering. What most of these schemes have in common is a

combination of laboratory and field measurements together with observation.

The laboratory measurements often include the determination of uniaxial

compressive strength together with other mechanical properties.

Chappell and Maurice (1980) have attempted to adapt the RMR rock mass

classification scheme for predicting the deformation of a rock mass beneath a

foundation. The scheme proposed by Chappell and Maurice is based upon

composite elastic theory which combine the intact and joint material

properties. The authors do not consider yielding of the rock mass, making the

scheme unsuitable for weak rocks such as chalk which exhibit such behaviour

under relatively low applied stresses. There appears to be no comprehensive

rock mass rating scheme suitable for predicting settlement on weak rocks in

the literature.

Some examples of engineering grade classifications are given by Knill and

Jones (1965). These classifications were developed specifically for studying

the geological conditions for dam foundations. In the case of the Roseires

dam the dominant feature controlling the assessment of foundation conditions

was weathering of the gneiss, hence the classification focused on this aspect.

In the case of the Latiyan dam it was the jointing and the presence of shale

layers that were considered to be important. Both these classifications were

developed to ensure efficient and rapid mapping of geological conditions

across each dam site. Both classifications are site specific, as is the case with

most engineering grade classifications.

Classic among engineering grade classifications was that developed by Ward

et al. (1968) for the chalk at Mundford, Norfolk. Table 2.3/5 shows the

original classification and subsequent additions of Wakeling (1970) (Grade VI

and correlation with SPT N value). The classification was developed in order

that the mass compressibility of the chalk measured at discrete points could

175

Table 2.3/5 Extended visual classification of the chalk (after Clayton, 1990a)

Grade Original Description (Ward et at., 1968, I - V, Wakeling, 1970, VI)

VI Extremely soft structureless chalk, containing small lumps of intact chalk.

V Structure less melange. Unweathered and partially-weathered angular chalk blocks and fragments set in a matrix of deeply-weathered remoulded chalk. Bedding and jointing are absent.

-"

" ~ IV Friable to rubbly chalk. Unweathered or partially-weathered chalk with bedding and jointing present. Joints and small fragments closely spaced, ranging from 10mm apart to about 60mm apart. Joints commonly open up to 20mm and infilled with weathered debris and small unweathered chalk fragments.

III Rubbly to blocky chalk. Unweathered medium to hard chalk with joints 60mm to 200mm apart. Joints open up to 3mm, sometimes with secondary staining and fragmentary infilling.

II Medium hard chalk with widely-spaced closed joints. Joints more than 200mm apart. When dug out for examination purposes this material does not pull away along joint faces but fractures irregularly.

Hard, brittle chalk with widely-spaced closed joints. Details as for Grade II but the chalk is harder.

SPTN

<8

8-15

15-20

20-25

25-35

>35

Identification factors nonnally used in practice

Structure Jointing/particle sizes

Bedding and behaviour dominated by jointing chalk fines absent.

Bedding and behaviour dominated by jointing intact chalk lumps absent

Bedding and Joints: 10-60mm spacing, jointing <20mm aperture with present debris in fill

Bedding and Joints: 60-200mm jointing spacing, <3mm aperture, present possible infill

Bedding and Joints: >200mm spacing, jointing Omm aperture present

Bedding and AsH jointing present

Identification factors not nonnally considered

Hardness/strength Weathering

extremely soft

deeply weathered

friable to rubbly unweathered or partially weathered

rubbly to blocky unweathered, sometimes with secondary joints

medium hard unweathered

hard unweathered

be extrapolated to other parts of the site which exhibited similar geological

features. Ward et al. (1968) assumed that the compressibility of the rock mass

depended primarily upon the following factors:

(i) The presence or absence of structure

(ii) Spacing of discontinuities

(iii) Orientation of discontinuities

(iv) Tightness (aperture) of discontinuities

(v) Hardness of intact chalk

The classification of the Mundford chalk was based on these factors. Grades

IV and V are largely the result of weathering (stress relief and frost

weathering) and are considered to be independent of lithology. Grades I and

IT are completely unweathered, the difference between them being a related

to lithological differences. A bed of hard low porosity chalk (Chalk Rock)

was encountered at depth in some parts of the site. It is this hard brittle

chalk that is classified as grade I.

In general grades V to ill occur in succession from the surface down

(sometimes with one or more missin~). Since grade I chalk relates to a

particular stratum it is possible for grade I chalk to overly grade II chalk.

Grade III chalk usually overlies grade II or I but occasionally it was found at

depth associated with minor tectonic features. The grades have been

correlated with rock mass properties such as compressibility and P-wave

velocity (Ward et al. 1968, Grainger et al., 1973). The correlations are shown

in Table 2.3/6. It can be seen from this table that the deformation

characteristics of chalk in the mass show a general improvement from grade

V through to I. Fig. 2.3/19 shows the range of values of Ei, Ey and qe

reported in the literature. The range of porosity involved in these

measurements is quite small, with a predominance in the more bighly

weathered chalks thought to have porosities of around 44%.

177

Table 2.3/6 Correlation of engineering grade with mechanical properties of

chalk in the mass.

Grade Approx range Yield Pwave Creep properties of E- stress velocity 1

'Ie Vp (MPa) (kPa) (mjs)

V 500 <200 650- 750 Exhibits significant creep IV 500-1000 200-400 1000-1200 as above

III 1000-2000 400-600 1600-1800 For pressures <400kPa creep is small and terminates in a few months

II 2000-5000 > 1000 2200-2300 Negligible creep for pressures at least 400kPa

I >5000 >1000 as above

(After Ward et aI (1968) & Grainger et aI. (1973))

Whilst structure, discontinuity spacing and aperture were defined in a way

that could be determined with relative ease in the field, the other parameters

were not. The lack of definition of some components may stem from the fact

that this classification was never intended to be used at sites other than

Mundford. Indeed Ward et al. (1968) state:

''It must be emphasised that this particular classification was developed

specifically for the conditions and requirements of the Mundford site. II

Despite this warning this classification scheme and its subsequent extensions

have been used by geotechnical engineers indiscriminately over much of the

chalk outcrop in the UK. The stratigraphic, pala:!ogeographic and tectonic

setting of the chalk at Mundford is now seen as atypical and hence the

correlations derived for this site cannot be applied to the outcrop as a whole.

Fookes and Horswill (1970) point out that the various components of the

classification at one grade at Mundford cannot necessarily be expected to be

found in unison at any other location because of the variable nature of this

rock.

178

In practice classification of chalk normally takes place on the basis of SPT N

value, the presence or absence of structure and discontinuity spacing. With

regard to the use of SPT N value to classify the chalk, the results are often

misleading. Dennehy (1975) compared the grade of chalk observed in trial

pits with that determined from SPT N values recorded in adjacent boreholes

formed by light percussion techniques. This study demonstrated that for a

given grade a wide range of SPT values could be obtained. Lord and Smith

(1976) carried out a similar study and concluded that SPT results do not

provide a suitable means of establishing chalk grade and hence foundation

level. Fig. 2.3/20 shows the results of standard penetration tests and visual

grading of the chalk for a silo complex at Bury St Edmunds. Based on the

SPT N values the chalk above the water table would be classified as Grade V

and yet the visual classification shows the quality of the chalk improving with

depth from Grade VI/V near the surface to Grade IV jm at depth. The

boundaries defined by the visual classification cannot be identified from the

SPT results. This example reinforces the point that SPT N values alone can

be misleading when attempting to determine the quality of the chalk for

foundation design.

The classification scheme proposed by Ward et al (1968) takes no account of

factors such as joint orientation, joint surface topography, the degree of

contact across joints and the looseness of the fracture block system. These are

significant factors which control the deformation of a rock mass and their

absence from this classification scheme serves to emphasise its site specific

nature.

179

Summary

• Theoretically the pressuremeter is the best method for measuring the

stiffness parameters for the rock mass. This is because in an elastic

continuum, all the elements of ground within the test zone will follow

similar stress paths. This means that stiffness parameters can be

determined without making any simplifying assumptions. Hence the

parameters may be used directly in predicting the settlement of a full

scale structure without the application of any scaling factors.

• the parameters E i, EY' qe and qy may be determined using the

pressuremeter.

• There are some serious practical disadvantages with the

pressuremeter. One of the principal disadvantages identified was that

the pressuremeter loads the ground in a different direction to a

foundation loading. The other important practical disadvantage with

pressuremeters was the difficulty in penetrating large flints.

• The plate loading test offers the best similitude between the test and

the full scale foundation. However the rigidity and geometry of the

plate results in different elements of the rock mass beneath the plate

following different stress paths. This results in problems extrapolating

from the test scale to that of a full-scale foundation.

• the parameters E i , EY' qe and qy may be determined using the plate

loading tests.

• In general the plate should have a diameter which is greater than five

times the average sub-horizontal discontinuity spacing in order to

minimise the scale effects and obtain reliable results.

180

• The interpretation of plate loading test data generally assumes the

ground to have a uniform stiffness. In non-homogeneous ground this

assumption imposes a serious limitation on interpretation.

• Values of Young's modulus are determined using pressure-settlement

data from plate loading tests. Following equations are used:

or

Rigid plate

Flexible plate

E = !:.qD (1 - v2)

4 P Pcentre

E = O.844qDp

(1 - v2

)

Pav

E = qD (1 - v 2)

p Pcentre

In practice stiffness parameters are determined assuming the plate to

be rigid.

• The rate of chage of stiffness with depth in non-homogeneous ground

may be determined from sub-plate settlement measurements or from

the distribution of ground surface settlement. Both methods are

generally unreliable in fractured chalk since they assume the ground to

be an elastic continuum.

• The major practical disadvantages of the plate loading test, particularly

when using large diameter plates, is that it is very expensive (generally

about £ 10,000 per test) to perform, it takes a long time to set up and

perform, it requires careful preparation of the ground and careful

selection and setting up of the instrumentation.

181

• Recent research work on the small strain stiffness behaviour of soils

there is a growing interest in the use of seismic methods to determine

stiffness parameters for soil and rock masses.

• Seismic techniques permit the measurement of stiffness that is close to

Ei since chalk does not exhibit significant non-linear stress-strain

behaviour at stresses less than qe.

• Seismic surveying techniques such as refraction, crosshole, uphole,

downhole and continuous surface wave permit a profile of modulus

with depth to be determined. This is clearly of importance in non­

homogeneous ground.

• P-waves may be used in structured chalk above and below the water

table. In saturated structureless chalk S-waves or Rayleigh waves must

be used since the P-wave velocity will not reflect the stiffness of the

material since it is dominated by the pore water.

• The advantage of geophysics is that it is relatively cheap and the

fieldwork can be conducted rapidly. The disadvantage is that the

geophysics is unable to provide data to enable the yield stress or post

yield modulus to be determined due to the very small strains induced

by seismic energy in the ground.

• The Standard Penetration Test does not measure the stiffness of the

rock mass. Values of stiffness can only be derived from empirical

correlations with SPT N value.

182

• Various empirical correlations between stiffness (Ei and Ey) and SPT

N value have been found. These include:

Wakeling (1966) Log linear relationship

Wake ling (1970) Log linear relationship

Lake and Simons (1970) Log linear relationship

Kee and Clapham (1971) Log linear relationship

Stroud (1988) Linear relationship

Powell et aL (1990) Log linear relationship

• The most commonly used empirical correlations in practice are

Wakeling (1970) and Kee and Clapham (1971).

• Most of the empirical relationships are unreliable since they are based

on a relatively small number of small-diameter plate tests. Of all the

relationships that proposed by Stroud (1988) appears to have the most

potentiaL

• In general stiffness values based on SPT N values are unreliable

principally because:

(i)

(ii)

SPT N value is sensitive to drilling method.

SPT N value is insensitive to relatively large changes in

rock mass stiffness.

• The Mundford engineering grade classification for chalk is site specific

but is applied by engineers across the whole of the chalk outcrop.

• The classification scheme is complicated by having too many classes

and having a number of descriptors which are not adequately defined

to be used in practice (eg weathering and hardness).

183

• In practice the classification of the chalk normally takes place on the

basis of:

SPT N value;

the presence or absence of structure

and discontinuity spacing.

• The use of SPT N values alone to classify the chalk is often

misleading.

184

Pressure volumeter.

Air to guard cells

Ground level

Preformed borehole

Probe.

Water injection

4+---Air

. .

...... --Water

C02 bottle

.--- Guard cell (Airfilled )

-..--- Measuring cell (Waterfilled)

4----- Guard cell ( Airfilled )

Fig. 2.3/1 Typical arrangement for a Menard pressuremeter test.

185

.-.. as D. :IE ---a. .. CD -::s tn tn CD -D.

Fig. 2:3/2

.-.. E ---.c ... a. CD C

Fig. 2.3/3

16

14

12

10

8

6

4

2

P f

p o

dp E = - (1 +v)

dEc - 1340 MP

Phase 3 - a

Phase 2

o 7~1I~~~~~~L---~--~ 10 11

0

2

4

6

Phase 1 Cavity strain, (%)

Typical curve from a Me d Ervin et al., 1980). nar pressuremeter test in moderately weak rock (after

o

Range of values back analysed from foundation settlements

74 .. .. •

20 30 40 50 60

Shear modulus, G (MPa)

Initial Reload loading cycle

Pressuremeter 0 •

Plate tests 6. £. 1710mm Dia. plate test *

.. 96 .. ......

70

Moduli determined from pressuremeter and plate tests on chalk compared with those back-analysed from a bridge foundation (from Marsland and Powell, 1983

and Powell et aI., 1990).

186

Borehole wall

Fig. 2.3/4

Loading column Locating cone

Fins

C=f=t~~~tlct==r=-~ Linear displacement transducers

Plaster

865mm Dia. plate

Plan

Settlement Elevation measurement ______ _

points

50mm

Springs

Settlement point

NOT TO SCALE

865mm diameter loading plate with four-point sub-plate ground-deformation measuring system (after Marsland and Eason, 1973).

187

Plate pressure (kPa)

o 200 400 600 800 0.0 r-----,-~~~----,_--~----_,----~----~----~

.-. 0.1 tft "-'" 0 --.. CG '-.. C 0.2 CD E CD -= CD en

0.3

Fig. 2.3/5

o

Plate

1

2

3

4

Sub-plate settlement measurement points set at 150mm intervals

"

COWDEN TILL

" " " " " " " " " " " "

" " " " " " "

'. '.

Plate

4

3

2

1

Typical load-settlement curves from sub-plate deformation measurements (after Hird et aI., 1991).

Error In E (%)

40 80 HID, = 10 0.0 ~---r-----r-----'--:-:-:-:-;-;-;-;----:r~----'

0.5

z/Dr

1.0

1.5

Fig. 2.3/6

.: I ~ \

" " ...... ""'-

" " '\ \ , , , , , ,

Load factor 0.15 0.25 0.33 0.50

H

j z

Plate Dia. = Dp Rigid plate assumed poisson's ratio = 0.5

Error in modulus values derived from sub-plate deformation measurements (after Hillier, 1991).

188

Fig. 23/7

E or G

0.0001

Field strains around structures ...

0.001 0.01 0.1 1

Normal or shear strain (%)

10

Conventional triaxial apparatus ... Local measurement of axial strain

I I

Resonant column

Geophysics ...

Typical relationship between stiffness and strain for soils.

189

Fig. 2.3/8

(a) Seismic refraction

Continuous increase in velocity with depth

H--------e.I-.

(b) Crosshole

Uphole Downhole

• • Energy source

• Detector (geophone)

(c) Uphole and downhole

Vibrator

(d) Surface wave

Seismic methods for determining the variation of stifness with depth in soils and rocks.

190

...... ~ ......

w

30

25

C 0 Q)

> .8 20 as t/) Q) '-1i)

:E 15

10 o

Fig. 2.3/9

E VI dl ng

SH1 T4

Vo= 250m/s 0

D1 D D2

V1 = 700m/s

V2 = 1000m/s IV

III III

V3 = 1600m/s ~

II V 4= 2300m/s _ .. _ .. _ .. - .. _ .. _ .. _ .. _ .. - .. _ .. _ .. _ .. - .. _. .._ .. _ ..

20 40 60 80 Metres

Estimated water table: March 1970, based on borehole Information

D1 Sandy soli D1 Cryoturbated bed

Velocity profile along seismic line 7 at Mundford, Norfolk (after Grainger et al. 1973).

50

CD E 30 0-... ... 0;; c 20 ca ~

10

• Survey measurements 4 1

Sinh fit v 0 = 377m/s, k = 100s

O~------~------~----__ L-______ L-____ ~ o

Fig. 2.3/10

u10 E as ~8

W'TJ ...

tn tn CD C = i

6

4

() 0- 2 E ca

10 20 30 40 50

Distance from energy source (m)

Typical time-distance graph from a seismic refraction test at Mundford, Norfolk . (after Abbiss, 1979).

Estatic= 0.47E dynamic

~ 0 ~~~ __ L-______ L-____ ~L-____ ~-------L------~ 6 C 0

Fig. 2.3/11

2 4

Static stiffness, E static (GPa)

Relationship betWeen static and dynamic moduli for the chalk at Mundford, Norfolk (after Abiss, 1979).

192

-. 5 E ---~ C ca ... ~ 10 -CD .a ~ ... D­CD Q 15

E (GPa) 4

• Seismic

• Finite element

.. Plate test

20 ~ __ -L ____ ~ __ ~ ____ ~ __ -L_

Fig. 2.3/12 Relationship between Young's modulus and depth for the Mundford chalk determined using seismic refraction, finite element analysis and plate loading tests (after Abbiss, 1979).

G (GPa) 00~ ______ 1~ ______ 2~ ______ 3~ ____ ~4

20

-. 40 E ---~ a CD 60 Q

80

I I I I I I

• ...

,

, \ ,

, , ,

Laboratory (static) tests Downhole

Cross hole

100 L-__ ---1 ___ .-.L ___ ---L----

Fig. 2.3/13 Shear modulus-depth profIles for chalk measured using cross-hole and down­hole seiesmic tests (after Sigismond et al., 1983).

193

10000 • Field observations

• Newbury lab. tests

1000 J E (MPa)

100

10

1

Fig. 2.3/14

10000

1000

E (MPa)

100

10

1

. Fig. 2.3/15

o

o

Norwich plate test

~ •

10

• • •

20

~--Medway Bridge Newbury pile test

30 40 50 60

SPT 'N' value

,

j

70

Correlation of stiffness with standard penetration test results for chalk (after Wakeling, 1966).

Line B ". :

.~

0.01 % of width

/ ..... 1-- Large settlements 0.5% of width

~ Medway bridge 9.S*32.3m

Newbury pile test 1372mm

66 Line A £. .... Purfleet plate tests 445mm Dia.

• Norwich plate test 140mm Dia.

~ Cambridge plate test 140mm Dia.

VI

Mundford Chalk Grades

10 20 30 40 50 60

SPT 'N' value

70

• Mundford tank • Mundford Plate (E.) 0 Mundford plate (E, ) • Other sites

Correlation of stiffness with standard penetration test results for chalk (after

Wakeling, 1970).

194

-" \.0 U1

100000 ,­,

-as D.. :E -tn ::J -::J "'C 0 :E

Co) .-... tn as -w

stroud (1988) 1S*N , .....

, ..... , ,

10000 I I J'" "d 7"

--...... -.: .. :>" ' ,

, , ,

.. , ;' ....... ~ .. ········.:..1 1000 I ., ." F 7' __ ~., ::: .. _ .. -.~

, ,

.--- ......

100 I L I l,J -'--',. ~ Ir"- ........ ' -1 .... - I I / () ____ ~n _ .. _.~ .....• . _ - 1

---Stroud r1988) 5.5*N

Lake & Simons (1970) 10 I _ -.:-:-:::':'''''1 r\.~e or. ,-",U:lpnilJll (l ~ I II

1 o

90)

10 20 30 40

SPT N Value

o Data from Powell et al. (1990)

50 60

Fig. 2.3/16 Comparison of empirical correlations between E and SPT 'N' value for the chalk.

70

50

40 E'

Nso 30

(MPa)

10

o

Fig. 2.3/17

--.

o

2

.5.. 4

= 6 ~ CD C 8

10

12

Fig. 2.3/18

0.1

* Ward et aI. (1968) • Burland et aL (1974)

fBB Hobbs et aI. (1979) t:j8 .(30 pie tests) • Wakeling (1966)

o Lake & Simons (1975) D Burland et at. (1969)

o Woodland et at. (1988)

- Fletcher et at. (1964)

.& • ~"""A .. ~ .... ~ .... .. .. 0 ...... - .. ----A. /\

----~-~------

0.2 0.3 0.4

Variation of E' /Nro with degree of loading for chalk (after Stroud, 1988).

SPT 'N' value

:~ :-: __ e .

• , .. ~ ••• • • l:. • • .. -. ::. .., .. -# •

• . : . I - .

100

e

• •

• •

Contractor C

180

• •

SPT 'N' value

• . . . . . .. ,:. .....

• ... . • •

~ -, :

180

Contractor 0

Comparison of the results of standard penetration tests carried out by different contractors at the same chalk site (after Mallard, 1977).

196

Fig. 2.3/19

104

I .. -;...~ I : I I : I I : I

3 I 10 Initial

I : I

I I : I I : I I : I

tangent ~-:_I i-;-:

: : :

modulus 2 I EI 10

(MPa)

10 II III IV V

104

Reported ranges of moduli and yield bearing pressure for in-situ chalk under foundation loading (after Clayton, 199Oa).

197

0 0

5

.-. E 10

----• ..I

• CJ 15 ~ 0 -CD .a

20 .z:. a CD c

25

30

35

Fig. 2.3/20

SPT 'N' value

10 20 30 40 50 , , . ' , YIN .: .. ~ , , V/IV : ~ ... , 0,

,0 .' ~.: 0·:

'~e: .e: IYIIII o ,-

~1. ' ~ '. ,. ~ : .. o:~r·: . ,

- ~ ,: Visual grading ~ , , • .. , , .. , • , ~8· .- • ~

8 e. e' • ••• 00 •• • • • , o 0' • • '. • • • • • r

• • ...-..-..-VI V IV III II I Wakeling's classification (1970)

S.I. of April 1972 S.I. of May 1981

Results of Standard Penetration Tests and visual grading of the chalk (after Burland and Bayliss, 1990).

198

2.4 Summary

The following points arise from this literature review:

• Most white chalks (ie Middle and Upper Chalk) exhibit a relatively

uniform chemical composition.

• Depositional, diagenetic and tectonic processes give rise to a wide

range of porosity (9% to 52%).

• Because of the uniform chemical composition, porosity is the principal

factor controlling the intact mechanical properties such as strength and

stiffness. It has been shown that the strength and stiffness of intact

chalk can be expected to vary through two orders of magnitude.

• Chalk exhibits yielding and in some cases collapse which is associated

with the breakdown of the bonded structure. The higher the porosity

of the chalk the more pronounced is the yield point and there is a

greater likelihood of collapse. Prior to yielding the rock material

behaves in a stiff and more or less elastic manner. After yield and

when the rock has become completely destructured it behaves as a

granular soil.

• The discontinuity pattern in the chalk is dominated by sub-horizontal

bedding discontinuities and at least two sets of sub-vertical joints

except in areas affected by intense tectonic activity. The areas affected

by intense tectonic activity account for only a small proportion of the

total outcrop area of the chalk in the UK. Hence most engineering

works will be on chalk with sub-horizontal and sub-vertical fracture.

199

• Weathering processes such as stress relief and frost action tend to

bring about an increase in frequency of discontinuities through the

development of new fractures. These fractures are not as persistent as

the primary discontinuities associated with diagenesis and tectonism.

They are however particularly abundant near the ground surface where

these weathering processes are most active.

• The general weathering profile for the chalk is characterised by

structureless chalk grading downwards into structured chalk with

discontinuity frequency and aperture reducing with depth. This profile

is largely the result of intense freezing during the Pleistocene.

• Dissolution has resulted in the widening of discontinuity apertures, and

the reduction in contact area across fractures through changes in

surface topography. There has been no study reported in the literature

concerning the rate and form of discontinuity dissolution.

• There has been no systematic study that examine the potential

relationship between the intact mechanical properties on the chalk and

the pattern of features within the weathering profile (weathering style)

for a given weathering regime. The only documented study on the

characterization of weathering style in the chalk was reported by Ward

et al. (1968).

• The introduction of discontinuities within a rock mass results in a

significant increase in compressibility.

• The stress distribution below a foundation on fractured rock is

controlled largely by the discontinuity geometry and hance cannot be

predicted using theories of continuum mechanics.

200

• The mass compressibility of most chalks will be dominated by the

normal stress-closure behaviour of sub-horizontal discontinuities. The

principal factors affecting the compressibility of such a rock mass when

subjected to a foundation loading include:

• stress-strain behaviour of the intact rock

• degree of contact across discontinuity walls ( contact area)

• discontinuity spacing in relation to the dimensions of the loaded

area

• discontinuity aperture

• discontinuity infill

• The available literature suggests that for a foundation placed on a

rock mass dominated vertical and horizontal discontinuities the load­

settlement curve should be concave such that stiffness increases with

increasing load. However chalk with similar discontinuity orientations

the load-settlement curve (obtained from plate loading tests) is convex

such that above a certain load the stiffness is significantly reduced.

This behaviour is associated with yielding of the rock mass. The

mechanisms causing the rock mass to yield are not understood.

• Burland and Lord (1970) suggested that the load-settlement behaviour

of a plate or foundation may be modelled using five parameters:-

E· - initial modulus 1

Ey = post yield modulus

qe = yield bearing pressure (based on the onset of yield)

qy = yield bearing pressure (based on the establishment of E)

• Only six well documented case records on the behaviour of

foundations on chalk could be found in the literature. Of these only

four deal with foundations on structured chalk. The reliability of some

of these case records is brought into question because of the

assumptions made concerning the contact stresses and the stress

201

distribution with depth (eg Bury St Edmunds sugar silos and the

Mundford tank loading test).

• The case records highlighted the following points:

• Generally pre-yield settlements are small with settlement ratios

less than 0.05% at average foundation pressures up to 200kPa.

• Little is known about the post-yield load-settlement behaviour.

• Little is known about the mechanisms causing yield and the

magnitude of the yield bearing pressure.

• Little is known about the time-settlement behaviour of

foundations on chalk.

• Little is known about the influence of intact mechanical

properties of chalk on the mass compressibility behaviour.

• Little is known about the influence of weathering styles on the

mass compressibility behaviour of chalk.

• Only 11 well documented case records of plate loading tests on chalk

could be found in the literature. These shed little light on the points

listed above since the majority of them employed a plate diameter that

was too small to give reliable results. Also in many cases only Ei could

be determined and in all but one case no attention was paid to the

time-settlement behaviour of the chalk.

• Stiffness parameters for use in foundation settlement predictions may

be measured using the following in-situ tests:

• • • • •

Pressuremeter

Plate loading tests

Surface or borehole seismic surveys

The Standard Penetration Test

Visual assessment

202

It is clear from this review of past literature that the mechanisms which

control the mass compressibility behaviour of chalk are not fully understood.

This is largely as a result of a distinct lack of well documented records of

load-settlement relationships for large scale loading tests and full scale

foundations on chalk. This lack of understanding precluded the general (non

site specific) characterization of chalk in the mass for the purpose of

predicting compressibility behaviour.

This thesis attempts to improve the fundamental understanding of some of

the factors controlling the mass compressibility of chalk by performing a

series of large diameter plate loading tests on weathered chalk at sites which

display similar discontinuity characteristics but different intact strength and

stiffness properties. The results of such tests will provide valuable data for a

currently deficient database on foundation behaviour on chalk as well as

permitting progress to be made in characterizing the chalk.

This thesis also attempts to evaluate some of the methods discussed above for

determining stiffness parameters for the chalk. In particular the standard

penetration test and surface wave geophysics will be used to predict the mass

compressibility of the chalk at the locations of the plate loading tests.

203